{ "0710/0710.5082.txt": { "abstract": "% context heading (optional) % {} leave it empty if necessary {} % aims heading (mandatory) {Photoionization models so far are unable to account for the high electron temperature $T_e$([\\ion{O}{iii}]) implied by the line intensity ratio [\\ion{O}{iii}]$\\lambda$4363\\AA/[\\ion{O}{iii}]$\\lambda$5007\\AA\\ in low-metallicity blue compact dwarf galaxies, casting doubts on the assumption of photoionization by hot stars as the dominant source of heating of the gas in these objects of large cosmological significance.} % methods heading (mandatory) {Combinations of runs of the 1-D photoionization code NEBU are used to explore alternative models for the prototype giant \\ion{H}{ii} region shell I\\,Zw\\,18\\,NW, with no reference to the filling factor concept and with due consideration for geometrical and stellar evolution constraints.} % results heading (mandatory) {Acceptable models for I\\,Zw\\,18\\,NW are obtained, which represent schematically an incomplete shell comprising radiation-bounded condensations embedded in a low-density matter-bounded diffuse medium. The thermal pressure contrast between gas components is about a factor 7. The diffuse phase can be in pressure balance with the hot superbubble fed by mechanical energy from the inner massive star cluster. The failure of previous modellings is ascribed to (1) the adoption of an inadequate small-scale gas density distribution, which proves critical when the collisional excitation of hydrogen contributes significantly to the cooling of the gas, and possibly (2) a too restrictive implementation of Wolf-Rayet stars in synthetic stellar cluster spectral energy distributions. A neutral gas component heated by soft X-rays, whose power is less than 1\\% of the star cluster luminosity and consistent with CHANDRA data, can explain the low-ionization fine-structure lines detected by SPITZER. [O/Fe] is slightly smaller in I\\,Zw\\,18\\,NW than in Galactic Halo stars of similar metallicity and [C/O] is correlatively large.} % conclusions heading (optional), leave it empty if necessary {Extra heating by, \\eg, dissipation of mechanical energy is not required to explain $T_e$([\\ion{O}{iii}]) in I\\,Zw\\,18. Important astrophysical developments are at stakes in the 5\\% uncertainty attached to \\oiii\\ collision strengths.} ", "introduction": "\\label{intro} The optical properties of Blue Compact Dwarf (BCD) galaxies are similar to those of Giant Extragalactic \\hii\\ Regions (GEHIIR). Their blue continuum arises from one or several young Massive Star Clusters (MSC), which harbour extremely large numbers of massive stars. BCDs are relatively isolated, small-sized, metal-poor galaxies (Kunth \\& \\\"Ostlin 2000) and may be the rare `living fossils' of a formerly common population. BCDs can provide invaluable pieces of information about the primordial abundance of helium (\\eg, Davidson \\& Kinman 1985), the chemical composition of the InterStellar Medium (ISM, \\eg, Izotov et al. 2006), the formation and evolution of massive stars, and the early evolution of galaxies at large redshift. Among them, \\IZ\\ stands out as one of the most oxygen-poor BCDs known (\\eg, Izotov et al. 1999) and a young galaxy candidate in the Local Universe (\\eg, Izotov \\& Thuan 2004). The line emission of \\hii\\ regions is believed to be governed by radiation from massive stars, but spectroscopic diagnostics most often indicate spatial fluctuations of the electron temperature \\Te\\ (see the dimensionless parameter $t^2$, Peimbert, 1967), that appear larger than those computed in {\\sl usual} photoionization models, suggesting an {\\sl extra heating} of the emitting gas (\\eg, Peimbert, 1995; Luridiana et al. 1999). Until the cause(s) of this failure of photoionization models can be identified, a sword of Damocles is hung over a basic tool of astrophysics. Tsamis \\& P\\'equignot (2005) showed that, in the GEHIIR 30\\,Dor of the LMC, the various \\Te\\ diagnostics could be made compatible with one another if the ionized gas were {\\sl chemically inhomogeneous} over small spatial scales. A pure photoionization model could then account for the spectrum of a bright filament of this nebula. Although this new model needs confirmation, it is in suggestive agreement with a scenario by Tenorio-Tagle (1996) of a recycling of supernova ejecta through a rain of metal-rich droplets cooling and condensing in the Galaxy halo, then falling back on to the Galactic disc and incorporating into the ISM without significant mixing until a new \\hii\\ region eventually forms. If this class of photoionization models is finally accepted, extra heating will not be required for objects like 30\\,Dor, with near Galactic metallicity. Another problem is encountered in low-metallicity (`low-Z') BCDs (Appendix~A). In BCDs, available spectroscopic data do not provide signatures for $t^2$'s, but a major concern of photoionization models is explaining the high temperature \\Te(\\oiii) infered from the observed intensity ratio $r$(\\oiii) = \\oiii\\la4363/(\\oiii\\la5007+4959). Thus, Stasi\\'nska \\& Schaerer (1999, SS99) conclude that photoionization by stars fails to explain $r$(\\oiii) in the GEHIIR \\IZ\\,NW and that photoionization must be supplemented by other heating mechanisms. A requirement for extra heating is indirectly stated by Luridiana et al. (1999) for NGC\\,2363. A possible heating mechanism is conversion of mechanical energy provided by stellar winds and supernovae, although a conclusion of Luridiana et al. (2001) does not invite to optimism. A limitation of this mechanism is that most of this mechanical energy is likely to dissipate in hot, steadily expanding superbubbles (Martin, 1996; Tenorio-Tagle et al. 2006). It is doubtful that heat conduction from this coronal gas could induce enough localized enhancement of \\Te\\ in the photoionized gas (\\eg, Maciejewski et al. 1996), even though Slavin et al. (1993) suggest that turbulent mixing may favour an energy transfer. Martin (1997) suggests that shocks could help to explain the trend of ionization throughout the diffuse interstellar gas of BCDs, but concedes that ``shocks are only being invoked as a secondary signal in gas with very low surface brightness''. Finally, photoelectric heating from dust is inefficient in metal-poor hot gas conditions (Bakes \\& Tielens 1994). Nevertheless, the conclusion of SS99 is now accepted in many studies of GEHIIRs. It entails so far-reaching consequences concerning the physics of galaxies at large redshifts as to deserve close scrutiny. If, for exemple, the difference between observed and computed \\Te(\\oiii) in the model by SS99 were to be accounted for by artificially raising the heat input proportionally to the photoionization heating, then the total heat input in the emitting gas should be doubled. {\\sl This problem therefore deals with the global energetics of the early universe}. After reviewing previous models for \\IZ\\,NW (Sect.~\\ref{prev_IZ}), observations and new photoionization models are described in Sects.~\\ref{obs} \\& \\ref{newmod}. Results presented in Sect.~\\ref{res} are discussed in Sect.~\\ref{disc}. Concluding remarks appear in Sect.~\\ref{concl}. Models for other GEHIIRs are reviewed in Appendix~A. Concepts undelying the new photoionization models are stated in Appendices~B and C. ", "conclusions": "\\label{concl} Owing to its small heavy element content, \\IZ\\ stands at the high-\\Te\\ boundary of photoionized nebulae. Where ionization and temperature are sufficiently high, the cooling is little dependent on conditions, except through the concentration of H$^0$, controlled by density. Therefore, in the photoionization model logics, {\\sl \\Te\\ is then a density indicator}, in the same way it is an O/H indicator in usual \\hii\\ regions. It is for not having recognized implications of this new logics, that low-metallicity BCD models failed. In a photoionization model study of \\IZ\\,NW, SS99 employed a filling-factor description and concluded that \\Te(\\oiii) was fundamentally unaccountable. The {\\sl vogue} for this simple description of the ionized gas distribution resides in its apparent success for usual \\hii\\ regions, a success falling in fact to the strong dependence of cooling on abundances. Universally adopted along past decades, this concept led all authors to conclude that photoionization by hot stars did not provide enough energy to low-Z GEHIIRs. This conclusion is in line with a movement of calling into question photoionization by stars as the overwhelmingly dominant source of heat and ionization in gaseous nebulae, a movement cristalizing on the `$t^2$ problem' (Esteban et al. 2002; Peimbert et al. 2004), since the presence of \\Te\\ fluctuations {\\sl supposedly larger} than those reachable assuming photoionization by stars implies additional heating. A conclusion of the present study is that the gas distribution is no less critical than the radiation source in determining the line spectrum of \\hii\\ regions. Assuming pure photoionization by stars, the remarkable piece of information carried by the large \\Te(\\oiii) of \\IZ\\,NW is that the mean density of the \\oiii\\ emitting region is much less than \\Ne(\\sii), a low \\Ne\\ confirmed by line ratios \\oiv\\la25.9$\\mu$/\\heii\\la4686 and \\feiii\\la4986/\\feiii\\la4658. \\IZ\\,NW models comprising a plausible SED and respecting geometrical constraints can closely match almost all observed lines from UV to IR, including the crucial \\oiii\\la4363 (\\siv\\la10.5$\\mu$ is a factor 2 off, however). Thus, extra heating by, \\eg, dissipation of mechanical energy in the photoionized gas of low-metallicity BCD galaxies like \\IZ\\ is {\\sl not} required to solve the `\\Te(\\oiii) problem'. Moreover, since low-ionization fine-structure lines can be explained by soft X-rays, (hydrodynamical) heating is {\\sl not even} required in warm \\hi\\ regions protected from ionization and heating by star radiation. As a final note, on close scrutiny, the solutions found here {\\sl tend to be just marginally consistent} with observed $r$(\\oiii). Given the claimed accuracy in the different fields of physics and astrophysics involved, postulating a mechanical source of heating is premature, whereas a {\\sl 2--3\\% upward correction} to the collision strength for transition O$^{2+}$($^3$P\\,--\\,$^1$S) at \\Te~$\\sim$~2$\\times$10$^4$\\,K is an alternative worth exploring by atomic physics. Yet, another possibility is a substantial increase of the distance to \\IZ. Owing to accurate spectroscopy and peculiar conditions in \\IZ, important astrophysical developments are at stakes in the 5\\% uncertainty attached to \\oiii\\ collision strengths. If photoionized nebulae are shaped by shocks and other hydrodynamic effects, this does not imply that the emission-line intensities are detectably influenced by the thermal energy deposited by these processes. Unravelling this extra thermal energy by means of spectroscopic diagnostics and models is an exciting prospect, whose success depends on a recognition of all resources of the photoionization paradigm. Adopting the view that photoionization by radiation from young hot stars, including WR stars, is the only excitation source of nebular spectra in BCD galaxies, yet without undue simplifications, may well be a way to help progresses in the mysteries of stellar evolution, stellar atmosphere structure, stellar supercluster properties, giant \\hii\\ region structure and, {\\sl at last}, possible extra sources of thermal energy in BCDs." }, "0710/0710.4400_arXiv.txt": { "abstract": "{} {Bright HNC 1--0 emission, rivalling that of HCN 1--0, has been found towards several Seyfert galaxies. This is unexpected since traditionally HNC is a tracer of cold (10 K) gas, and the molecular gas of luminous galaxies like Seyferts is thought to have bulk kinetic temperatures surpassing 50 K. There are four possible explanations for the bright HNC: (a) Large masses of hidden cold gas; (b) chemistry dominated by ion-neutral reactions; (c) chemistry dominated by X-ray radiation; and (d) HNC enhanced through mid-IR pumping. In this work we aim to distinguish the cause of the bright HNC and to model the physical conditions of the HNC and HCN emitting gas.} {We have used SEST, JCMT and IRAM 30m telescopes to observe HNC 3--2 and HCN 3--2 line emission in a selection of 5 HNC-luminous Seyfert galaxies. We estimate and discuss the excitation conditions of HCN and HNC in NGC~1068, NGC~3079, NGC~2623 and NGC~7469, based on the observed 3--2/1--0 line intensity ratios. We also observed CN 1--0 and 2--1 emission and discuss its role in photon and X-ray dominated regions.} {HNC 3--2 was detected in 3 galaxies (NGC~3079, NGC~1068 and NGC~2623). Not detected in NGC~7469. HCN 3--2 was detected in NGC~3079, NGC~1068 and NGC~1365, it was not detected in NGC~2623. The HCN 3--2/1--0 ratio is lower than 0.3 only in NGC~3079, whereas the HNC 3--2/1--0 ratio is larger than 0.3 only in NGC~2623. The HCN/HNC 1--0 and 3--2 line ratios are larger than unity in all the galaxies. The HCN/HNC 3--2 line ratio is lower than unity only in NGC~2623, which makes it comparable to galaxies like Arp~220, Mrk~231 and NGC~4418. } {We conclude that in three of the galaxies the HNC emissions emerge from gas of densities $n\\lesssim10^5~\\3cm$, where the chemistry is dominated by ion-neutral reactions. The line shapes observed in NGC~1365 and NGC~3079 show that these galaxies have no circumnuclear disk. In NGC~1068 the emission of HNC emerges from lower ($<10^5~\\3cm$) density gas than HCN ($>10^5~\\3cm$). Instead, we conclude that the emissions of HNC and HCN emerge from the same gas in NGC~3079. The observed HCN/HNC and CN/HCN line ratios favor a PDR scenario, rather than an XDR one, which is consistent with previous indications of a starburst component in the central regions of these galaxies. However, the $N({\\rm HNC})/N({\\rm HCN})$ column density ratios obtained for NGC~3079 can be found only in XDR environments.} ", "introduction": "The hydrogene cyanide, HCN molecule, is commonly used as an extragalactic tracer of molecular gas with densities $n$(H$_2$) larger than $10^4~\\3cm$ (e.g. Solomon et al. \\cite{solomon92}; Curran et al. \\cite{curran00}; Kohno \\cite{kohno05}). The HCN to CO intensity ratio varies significantly, from 1/3 to 1/40 in starburst galaxies, and it has not been determined whether this variation depends on the dense molecular gas content or on the abundance and/or excitation conditions. In addition, recent results seems to indicate that HCN may not be an unbiased tracer of the dense molecular gas content in LIRGs and ULIRGs (Graci\\'a-Carpio et al. \\cite{gracia06}). It is essential, therefore, to use other molecular tracers than HCN, in order to understand the physical conditions in the dense gas. A molecule of particular interest, for comparison with HCN, is its isomer HNC. The detection of interstellar HNC supports the theory of dominant ion-molecule chemistry in dark molecular clouds. Both species are thought to be created by the same dissociative recombination of HCNH$^+$. This ion can produce HCN and HNC, with approximately equal abundances. Models based only on this scheme would predict then an HNC/HCN ratio $\\approx 1$. However, the observed HNC/HCN abundance ratios vary significantly between different kinds of molecular clouds - the ratio ranges from 0.03 to 0.4 in warm cores ($T_k > 15$ K), and can be as high as 4.4 in cold cores ($T_k < 15$ K). The CN (cyanogen radical) molecule is another tracer of dense gas, with a lower (by a factor of 5) critical density than HCN. CN is also chemically linked to HCN and HNC by photodissociations (e.g. Hirota \\& Yamamoto \\cite{hirota99}). Surveys of the 1--0 transition of CN and HNC have been done in order to trace a cold, dense phase of the gas in luminous galaxies (Aalto et al. \\cite{aalto02}). It was found that the HNC 1--0 luminosities often rivalled those of HCN 1--0. These results seem to contradict the idea of warm ($T_k \\gtrsim 50$ K) gas in the centers of luminous galaxies (e.g. Wild et al. \\cite{wild92}; Wall et al. \\cite{wall93}) whose IR luminosities were suggested to originate from star formation rather than AGN activity (Solomon et al. \\cite{solomon92}). \\begin{table*}[!t] \\begin{minipage}{18cm} \\caption[]{Sample of galaxies$^{~\\rm a}$.} \\label{tab:galaxies} \\centering \\begin{tabular}{lccccccc} \\hline \\noalign{\\smallskip} Galaxy & Seyfert & RA & DEC & v$_{sys}$ & Distance$^{~\\rm b}$ & $\\Omega_S$(CO)$^{~\\rm c}$ & $\\Omega_S$(HCN)$^{~\\rm d}$\\\\ & & [hh mm ss] & [$^{\\circ}~{}'~{}''$] & [\\kms] & [Mpc] & [$''~^2$] & [$''~^2$]\\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} NGC~3079 & 2 & 10 01 57.805 & +55 40 47.20 & 1116$\\pm$1 & 15.0$\\pm$1.1 & $15\\times 7.5$ & $5\\times 5$ \\\\ NGC~1068 & 2 & 02 42 40.711 & -00 00 47.81 & 1137$\\pm$3 & 15.3$\\pm$1.1 & $30\\times 30$ & $10\\times 10$ \\\\ NGC~2623 & 2 & 08 38 24.090 & +25 45 16.80 & 5549$\\pm$1 & 74.6$\\pm$5.4 & $8\\times 8$ & $2.6\\times 2.6$ \\\\ NGC~1365 & 1.8 & 03 33 36.371 & -36 08 25.45 & 1636$\\pm$1 & 22.0$\\pm$1.6 & $50\\times 50$ & $16.5\\times 16.5$ \\\\ NGC~7469 & 1.2 & 23 03 15.623 & +08 52 26.39 & 4892$\\pm$2 & 65.8$\\pm$4.8 & $8\\times 8$ & $4\\times 6$ \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\begin{list}{}{} \\item[${\\mathrm{a}}$)] The Seyfert classification, positions (in equatorial J2000 coordinates) and heliocentric radial velocities were taken from NED. \\item[${\\mathrm{b}}$)] The distances were calculated using the Hubble constant (H$_0 \\approx 74.37$ \\kms~Mpc$^{-1}$) estimated by Ngeow and Kanbur (\\cite{ngeow06}). \\item[${\\mathrm{c}}$)] The source sizes of the CO 1--0 transition line were estimated from the maps presented in (Koda et al. \\cite{koda02}) for NGC~3079, (Schinnerer et al. \\cite{schinnerer00}) for NGC~1068, (Bryant et al. \\cite{bryant99}) for NGC~2623, (Sandqvist \\cite{sandqvist99}) for NGC~1365, and (Papadopoulos \\& Allen \\cite{papadopoulos00}) for NGC~7469. The source sizes for the $J = 2-1$ transition line were assumed equal to that of the $J = 1-0$ line. For NGC~7469, the source size of the CO 2--1 emission estimated from the corresponding map presented by Davies et al. (\\cite{davies04}) agrees well with the source size estimated for the CO 1--0 line. \\item[${\\mathrm{d}}$)] Source sizes of HCN 1--0 were estimated from the corresponding maps published in (Kohno et al. \\cite{kohno00}) for NGC~3079, (Kohno et al. \\cite{kohno01} and Helfer \\& Blitz \\cite{helfer95}) for NGC~1068, (Davies et al. \\cite{davies04}) for NGC~7469. The source sizes of NGC~2623 and NGC~1365 were estimated using proportions found in NGC~1068 (read text in \\S3.5). Because of their chemical link, the source sizes of the CN and HNC molecules were assumed the same as that of HCN. Due to the lack of high resolution maps, the source sizes corresponding to the emission of the higher transition lines were assumed equal to that of the $J = 1-0$ line. \\end{list} \\end{minipage} \\end{table*} According to observations in the vicinity of the hot core of Orion KL, experimental data, chemical steady state and shock models, the HNC/HCN ratio decreases as the temperature and density increase (e.g. Schilke et al. \\cite{schilke92}; Talbi et al. \\cite{talbi96}; Tachikawa et al. \\cite{tachikawa03}). If a bright HNC 1--0 transition line is nevertheless detected under these conditions, it could be due to the following possible explanations: (a) the presence of large masses of hidden cold gas and dust at high densities ($n > 10^5~\\3cm$); (b) chemistry dominated by ion-molecule reactions with HCNH$^+$ at low density ($n \\approx 10^4~\\3cm$) in regions where the temperature dependence of the HNC abundance becomes weaker; (c) enhancement by mid-IR pumping, also in low density regions where the lines would not be collisionally excited; and (d) the influence of UV-rays in Photon Dominated Regions (PDRs) and/or X-rays in X-ray Dominated Regions (XDRs) at densities $n \\gtrsim 10^4~\\3cm$ and at total column densities $N_{\\rm H} > 3\\times 10^{21}~\\2cm$ (Meijerink \\& Spaans \\cite{meijerink05}). In the case of CN, observations of its emission towards the Orion A molecular complex (Rodr\\'iguez-Franco et al. \\cite{rodriguez98}) suggest that this molecule is also enhanced in PDRs, but particularly in XDRs, where a CN/HCN abundance ratio larger than unity is expected (e.g., Lepp \\& Dalgarno \\cite{lepp96} and Meijerink, Spaans, Israel \\cite{meijerink07}). We have observed low and high transition lines of the HCN, HNC and CN molecules in a group of Seyfert galaxies, which are supposed to host both sources of power, AGN and starburst activity, in their central region. Our interest is to assess the excitation conditions of HCN and HNC, distinguish between the above possible causes of the bright HNC, and to explore the relation between the CN emission, XDRs and dense PDRs in these sources. In \\S2 we describe the observations. The results (spectral lines, line intensities and line ratios) are presented in \\S3. The interpretation of line shapes and gas distribution in the most relevant cases, as well as the possible explanations for the bright HNC and the modelling of the excitation conditions of HCN and HNC are discussed in \\S4. The conclusions and final remarks of this work are presented in \\S5. ", "conclusions": "We have used the SEST and JCMT telescopes to carry out a survey of CN 2--1, HCN 3--2 and HNC 3--2 line emission in a sample of 4 Seyfert galaxies, plus NGC~3079 which was observed with the IRAM 30m telescope. The conclusions we draw are as follows: \\begin{list}{}{} \\item[1)] We detected HNC 3--2 emission in 3 of the 5 galaxies, while we obtain an upper limit for one of them (NGC~7469). HCN 3--2 was also detected in 3 galaxies (NGC~3079, NGC~1068 and NGC~1365), while it was not detected in NGC~2623. CN 2--1, along with the spingroups (\\textit{J} = 5/2 -- 3/2, \\textit{F} = 7/2 -- 5/2) and (\\textit{J} = 3/2 -- 1/2, \\textit{F} = 5/2 -- 3/2) was also detected in NGC~3079, NGC~1068 and NGC~1365. \\item[2)] {The line shapes} observed in NGC~1365 and NGC~3079 suggests that there is no circumnuclear disk in these galaxies. \\item[3)] We find that in 3 of the galaxies the HNC 3--2/1--0 line ratios suggest that the HNC emissions emerge from gas of densities $n\\lesssim10^5~\\3cm$, where the chemistry is dominated by ion-neutral reactions. In NGC~2623 a model of large masses of hidden cold (10 K) gas and dust, as well as a chemistry dominated by ion-neutral reactions, are yet to be distinguished as the correct interpretation for the bright HNC observed in this galaxy. \\item[4)] The 3--2/1--0 line ratios and the modelled excitation conditions imply that the HNC emission emerges from a more diffuse ($n<10^5~\\3cm$) gas region than the HCN emission ($n>10^5~\\3cm$) in NGC~1068, whereas they emerge from the same lower density ($n\\lesssim10^5~\\3cm$) gas in NGC~3079. \\item[5)] The HCN/HNC and CN/HCN line ratios tentatively favor a PDR scenario, rather than an XDR one, in the 3 Seyfert galaxies where we have CN, HNC and HCN data. The $N({\\rm HNC})/N({\\rm HCN})$ column density ratios obtained for NGC~3079 can be found only in XDR environments. \\end{list} In order to complete the sample, we plan to observe HCN 3--2 and CN 2--1 in NGC~7469, CN 2--1 in NGC~2623 and HNC 3--2 in NGC~1365. We plan to perform high resolution observations to further study the distribution and source sizes of CN and HNC. Modeling of the collision data for the CN molecule would be useful to estimate the $N({\\rm CN})/N({\\rm HCN})$ column density ratio, which would complement the $N({\\rm HNC})/N({\\rm HCN})$ ratio in order to have a more sophisticated tool to estimate and distinguish the prevalent environment conditions of the high density gas in the nuclear region of Seyfert galaxies. The AGN contribution (through XDR effects) is typically of a small angular scale and can be seriously affected by beam dilution at the transition lines studied in this work. On the other hand, the starburst contribution is of a larger angular scale than the AGN, and it effects can be contaminating our observations, and hence leading to the favored PDR scenario found with our models. Hence, our suggested interpretations could change if we zoom in on these sources. Therefore, high resolution maps of HNC and CN molecules are necessary to complement those of HCN, and to do a more accurate estimate of molecular abundances and line intensity ratios, which take source size into account. Observations of the higher transition lines (e.g. $J$=4--3) can also aid to disentangle the effects of the AGN and the starburst ring, due to the smaller beam size obtained at higher frequencies." }, "0710/0710.3685_arXiv.txt": { "abstract": "{ Relic neutralinos produced after the Big Bang are favoured candidates for Dark Matter. They can accumulate at the centre of massive celestial objects like our Sun. Their annihilation can result in a high-energy neutrino flux that could be detectable as a localised emission with earth-based neutrino telescopes like ANTARES. In this paper a brief overview of the prospects of the indirect search for Dark Matter particles with the ANTARES detector will be given. The analysis method and expected performance for the detection of the expected neutrinos will be discussed. \\PACS{ {95.35.+d}{Dark matter (stellar, interstellar, galactic, and cosmological)} \\and {95.55.Vj}{Neutrino, muon, pion, and other elementary particle detectors; cosmic ray detectors} } % } % ", "introduction": "\\label{intro} It is now a well-established fact that according to our present theory of gravity, 85\\%~of the matter content of our universe is missing. Observational evidence for this discrepancy ranges from macroscopic to microscopic scales, e.g. gravitational lensing in galaxy clusters, galactic rotation curves and fluctuations measured in the Cosmic Microwave Background. This has resulted in the hypothesised existence of a new type of matter called Dark Matter. Particle physics provides a well-motivated explanation for this hypothesis: The existence of (until now unobserved) massive weakly interacting particles (WIMPs). A favorite amongst the several WIMP candidates is the neutralino, the lightest particle predicted by Supersymmetry, itself a well-motivated extension of the Standard Model. If Supersymmetry is indeed realised in Nature, Supersymmetric particles would have been copiously produced at the start of our Universe in the Big Bang. Initially these particles would have been in thermal equilibrium. After the temperature of the Universe dropped below the neutralino mass, the neutralino number density would have decreased exponentially. Eventually the expansion rate of the Universe would have overcome the neutralino annihilation rate, resulting in a neutralino density in our Universe today similar to the cosmic microwave background. These relic neutralinos could then accumulate in massive celestial bodies in our Universe like our Sun through weak interactions with normal matter and gravity. Over time the neutralino density in the core of the object would increase considerably, thereby increasing the local neutralino annihilation probability. In the annihilation process new particles would be created, amongst which neutrinos. This neutrino flux could be detectable as a localised emission with earth-based neutrino telescopes like ANTARES. This paper gives a brief overview of the prospects for the detection of neutrinos originating from neutralino annihilation in the Sun with the ANTARES neutrino telescope. \\begin{figure}[b] \\center{ \\includegraphics[width=0.45\\textwidth,angle=0]{NEA_60kHz0XOFF_off.png} \\caption{The ANTARES Neutrino Effective Area vs. $E_\\nu$.} \\label{fig:1} % } \\end{figure} \\begin{figure*}[t] \\center{ \\includegraphics[width=0.8\\textwidth,angle=0]{psflux.png} \\caption{Predicted $\\nu_\\mu+\\bar{\\nu}_\\mu$ flux from the Sun in mSUGRA parameter space.} \\label{fig:2} % } \\end{figure*} ", "conclusions": "Nearly half of the ANTARES detector has been operational since January this year. The detector is foreseen to be completed in early 2008. The data show that the detector is working within the design specifications. As can be seen from Fig.~\\ref{fig:4}, mSUGRA models that are excludable by ANTARES at 90\\%~CL are found in the Focus Point region. The same models should also be excludable by future direct detection experiments, as is shown in Fig.~\\ref{fig:7}. To improve the ANTARES sensitivity, a directional trigger algorithm has recently been implemented in the data acquisition system. In this algorithm, the known position of the potential neutrino source is used to lower the trigger condition. This increases the trigger efficiency, resulting in a larger $A_{\\rm eff}^{\\nu}$. In Fig.~\\ref{fig:8}, the $A_{\\rm eff}^{\\nu}$ at the trigger level for the standard- and the directional trigger algorithm are shown in black (``{\\em trigger3D}'') and red (``{\\em triggerMX}'') respectively." }, "0710/0710.2360_arXiv.txt": { "abstract": "{ The similarity of the observed mass densities of baryons and cold dark matter may be a sign they have a related origin. The baryon-to-dark matter ratio can be understood in the MSSM with right-handed (RH) neutrinos if CDM is due to a d = 4 flat direction condensate of very weakly coupled RH sneutrino LSPs and the baryon asymmetry is generated by Affleck-Dine leptogenesis along a d = 4 $\\left(H_{u}L\\right)^2$ flat direction. Observable signatures of the model include CDM and baryon isocurvature perturbations and a distinctive long-lived NLSP phenomenology. \\PACS{ {98.80.Cq}{Cosmology} } % } % ", "introduction": "\\label{intro} A striking feature of the observed Universe is the similar mass density in baryons and cold dark matter. (The 'Baryon-to-Dark Matter' (BDM) ratio.) From the WMAP three-year results for the $\\Lambda$CDM model, $\\Omega_{DM}/\\Omega_{B} = 5.65 \\pm 0.58$ \\cite{wmap}. However, in most models the physics of baryogenesis and dark matter production are unrelated. So why is the mass density in baryons within an order of magnitude of that of dark matter? There are three possibilities: \\newline (i) A remarkable coincidence. \\newline (ii) Some anthropic selection mechanism, usually assumed but undefined (e.g. in the case of thermal relic neutralino dark matter). \\newline (iii) The mechanisms for the origin of the baryon asymmetry and dark matter are related. The latter possibility seems the simplest interpretation of the BDM ratio. Indeed, we may be ignoring a {\\it big clue} to the nature of the correct BSM particle theory. It is highly non-trivial for a particle physics theory to have within its structure (without contrivance) a mechanism that can account for the BDM ratio. Therefore if the BDM ratio is due to such a mechanism it would provide us with a powerful principle by which to select the best canadiate particle physics models. BDM models broadly divide into two classes: \\newline {\\bf 1).Charge conservation based:} The dark matter particle and baryon number are related by a conserved charge, $Q_{B} + Q_{cdm} = 0 \\Rightarrow n_{cdm} \\sim n_{B}$. The CDM particle mass satisfies $ m_{cdm} = m_{n}n_{B}/n_{cdm}$ and so $m_{cdm} \\sim 1 GeV$ is necessary. However, this does fit well with SUSY if the LSP mass is $O(m_{W})$ or larger. \\newline {\\bf 2). Dynamics based:} In this case the dark matter and baryon densities are related by {\\it similar} physical mechanisms for their origin. This implies a less rigid relation between $n_{B}$ and $n_{CDM}$, which may allow us to understand why it is the mass rather than number densities that are observed to be similar. ", "conclusions": "RH sneutrino condensate CDM combined with Affleck-Dine baryogenesis can plausibly account for the observed similarity of the baryon and dark matter mass densities in the Universe. Seen as a selection principle, the requirement that a particle physics model can {\\it without contrivance} account for the BDM ratio favours the MSSM with neutrino masses and RH sneutrino condensate CDM. It is quite remarkable that the MSSM with neutrino masses has the ability to account for the BDM ratio as a natural consequence its structure. CDM and baryon isocurvature perturbations are possible, the ratio of which gives information on the nature of SUSY inflation. Long-lived NLSP phenomenology is also expected, which can be distinguished from gravitino and axino LSP phenomenology via trapped stau final states, and from thermal relic RH sneutrino LSP phenomenology once the parameters of the MSSM are known. There are a number of issues which remain to be addressed. One is the possibility that $M_{N} \\neq 0$. In this case it is possible that the heavier generation condensate RH sneutrinos, which will have a lifetime longer than the age of the Universe so long as $\\lambda_{\\nu}$ is still small, would decay into the LSP RH sneutrino plus $e^{+}e^{-}$. This could produce a potentially observable diffuse $\\gamma$-ray background. In addition, the phenomenology and cosmology of MSSM-LSPs in this model should be studied in detail." }, "0710/0710.2683_arXiv.txt": { "abstract": "We study the dynamics of a twisted tilted disc under the influence of an external radiation field. Assuming the effect of absorption and reemission/scattering is that a pressure is applied to the disc surface where the local optical depth is of order unity, we determine the response of the vertical structure and the influence it has on the possibility of instability to warping. We derive a pair of equations describing the evolution of a small tilt as a function of radius in the small amplitude regime that applies to both the diffusive and bending wave regimes. We also study the non linear vertical response of the disc numerically using an analogous one dimensional slab model. For global warps, we find that in order for the disc vertical structure to respond as a quasi uniform shift or tilt, as has been assumed in previous work, the product of the ratio of the external radiation momentum flux to the local disc mid plane pressure, where it is absorbed, with the disc aspect ratio should be significantly less than unity. Namely, this quantity should be of the order of or smaller than the ratio of the disc gas density corresponding to the layer intercepting radiation to the mid plane density, $\\lambda \\ll 1$. When this condition is not satisfied the disc surface tends to adjust so that the local normal becomes perpendicular to the radiation propagation direction. In this case dynamical quantities determined by the disc twist and warp tend to oscillate with a large characteristic period $T_{*}\\sim \\lambda^{-1}T_{K}$, where $T_{K}$ is some 'typical' orbital period of a gas element in the disc. The possibility of warping instability then becomes significantly reduced. In addition, when the vertical response is non uniform, the possible production of shocks may lead to an important dissipation mechanism. ", "introduction": "\\noindent A significant number of X-ray binaries exhibit long-term periodicities on time-scales of ~10-100 d. Examples are Her X-1, SS 433 and LMC X-4 see e.g. Clarkson et al. 2003. Precession and warping of a tilted accretion disc has been proposed as an explanation (Katz 1973; Petterson 1975) which has subsequently been found to have observational support (Clarkson et al. 2003). The effect of radiation pressure on a twisted tilted accretion disc was first considered by Petterson 1977a,b who noted that when radiation from the central source is absorbed at the disc surface and re-emitted, a potentially important torque will result. Iping \\& Petterson 1990 later suggested that such torques determined the shape of the disc and its precession rate. Pringle 1996 subsequently showed that an initially axisymmetric thin disc could be unstable to warping as a result of interaction with an external radiation field (see e.g. Maloney et a. 1998; Ogilvie \\& Dubus 2001 and Foulkes et al. 2006 for later developments). The analysis considered the disc to behave as a collection of rings, which could interact by transferring angular momentum through the action of viscosity, but otherwise behaved as if they were rigid. In particular, the vertical displacement or tilt was assumed to be essentially uniform and independent of the vertical coordinate. Thus the pressure applied at the surface at optical depth unity is assumed to be effectively communicated through the disc vertical structure so as to give a near uniform response. In this paper we extend the theoretical treatment of the interaction of a disc with an external radiation field to take account of possible significant departures of the tilt or displacement from uniformity in the vertical direction. One of our objectives is to determine the conditions under which the assumption of a uniform response is valid, and then to estimate some of the consequences when they are not satisfied, including potential additional dissipation resulting from non linear effects such as the production of shock waves. The latter is done using a one dimensional slab analogue model with the required high resolution in the vertical direction. Although we focus on the effects of a surface pressure induced by an external radiation field, very similar considerations are likely to apply when warps are induced by a surface pressure resulting from interaction with an external wind (e.g. Quillen, 2001) or surface forces produced by the interaction of the disc with an external magnetic field originating in the central star ( e.g. Pfeiffer \\& Lai, 2004). We begin by considering the relevant issues using simple physical arguments. Following Pringle 1996 we consider a thin disc immersed in an external radiation field with mid plane initially coinciding with a Cartesian $(x,y)$ plane. It is supposed that the upper surface is parallel to the external rays so that there is initially no interaction with the radiation. The upper surface is then given a vertical elevation $h(r,\\phi),$ where we now use polar coordinates $(r,\\phi).$ As a result, a disc element with surface area $d{\\cal A}$ absorbs momentum from the radiation field at a rate \\be {\\dot {\\cal F}} = F_0 d{\\cal A} \\left(r{\\partial (h/r)\\over \\partial r}\\right).\\ee Here $ F_0$ is the momentum flux at radius $r$ associated with the external radiation field. The factor in brackets gives the angle through which the local normal is rotated as a result of the perturbation (see Appendix \\ref{A} for more details). Assuming the absorbed momentum is reradiated isotropically above the disc by the surface layers at an optical depth of unity, there will be an applied pressure there of magnitude \\be {\\cal P} = {2F_0\\over 3} \\left(r{\\partial (h/r)\\over \\partial r}\\right).\\ee This external pressure, when applied to a complete elementary ring, produces a net torque per unit length of magnitude \\be { d {\\cal T} \\over dr} = {2\\pi F_0 \\over 3} \\left(r^2{\\partial (h/r)\\over \\partial r}\\right){\\bf l},\\ee where the complex vector ${\\bf l}$ has components equal to $(i,1,0)$ in the Cartesian coordinate system. Here, in performing the azimuthal integration, we take into account that the azimuthal dependence of $h$ is through a factor of the form $ \\exp(-i\\phi)$ and work with the radial amplitude from now on. To find the consequent evolution of the disc, one requires the component of the angular momentum per unit length perpendicular to its unperturbed direction. This is given by \\be {d{\\cal J}/dr} = 2\\pi\\Sigma r^2 \\Omega {\\langle h \\rangle}i{\\bf l},\\ee with $\\Omega$ and $\\Sigma $ being the near Keplerian local disc angular velocity and the disc surface density, respectively. $\\langle h \\rangle =\\int d\\zeta \\rho h /\\Sigma $ where $\\zeta $ is a vertical coordinate and $\\rho $ is the gas density. For an elementary ring the condition that the rate of change of angular momentum equal the applied torque gives a tilt evolution equation of the form \\be {\\partial {\\langle h \\rangle}\\over \\partial t}= -i{ F_0 \\over 3\\Omega\\Sigma} \\left({\\partial (h/r)\\over \\partial r}\\right). \\ee Assuming the vertical response is uniform, so that we can set $h= {\\langle h \\rangle} \\equiv r{\\bf W}$, we obtain a description of warp evolution equivalent to Pringle 1996 when effects due to viscosity and bending wave propagation are neglected (see sections \\ref{simpa} and \\ref{Dyneq} below). This description indicates the possibility of instability to radiation pressure warping. However, here we stress the fact that the averaged elevation $\\langle h \\rangle ,$ enclosed in angled brackets, applies to the bulk of the inertia of the disc and can therefore be shown to be close to the elevation of the mid plane. On the other hand $h$ as used in the torque formula applies to the disc surface elevation. An important aspect of this paper is to distinguish these two elevations and investigate under what conditions they can be taken to be equal. This would be possible if the vertical structure responds as a rigid body to the external pressure forcing and we find the conditions for this to occur. The general requirement is found to be that the density at the surface of the disc where the pressure is applied should not be too small. It is possible to estimate in a simple way when the vertical displacement response of the disc to the external pressure forcing becomes non uniform. Let us suppose that the external pressure $ {\\cal P}$ is applied at the surface layer where the density $\\rho = \\rho_*.$ The induced vertical displacement $h,$ now should be considered to be also a function of the vertical coordinate $\\zeta.$ Assuming a linear response, which should be appropriate for sufficiently small elevations and, for simplicity an isothermal equation of state with sound speed $c_s,$ it can be easily shown that in the upper layers of the disc where vertical motions dominate the Lagrangian pressure perturbation has the form determined by presence of $h$, $\\Delta P = -\\rho_{*}c_{s}^{2} {\\partial h\\over \\partial \\zeta}$, see equation (\\ref{e24}) below. Equating this to the external pressure we have \\be {\\partial h\\over \\partial \\zeta}= -{{\\cal P}\\over \\rho_* c_s^2}.\\ee If the vertical extent of the disc is $\\zeta_0,$ and $h$ varies on a radial scale comparable to $r,$ the characteristic change in $h$ induced over the vertical thickness is easily estimated to be \\be \\Delta h \\sim \\zeta_0{{\\cal P}\\over \\rho_* c_s^2} \\sim {2F_0r \\over 3\\rho_* \\Omega^2\\zeta_0} \\left({\\partial (h/r)\\over \\partial r}\\right),\\label{i1}\\ee where we use the approximate relation $c_s=\\Omega\\zeta_0.$ From the condition $\\Delta h \\sim h $, using (\\ref{i1}) and estimating ${\\partial h\\over \\partial r}\\sim h/r$, it is clear that whether $h$ is uniform or not is governed by the the parameter $ F_0 r /(\\rho_* \\Omega^2\\zeta_0^3)(\\zeta_0/r)^2 = \\lambda^{-1} \\epsilon \\delta^2,$ with $\\lambda =\\rho_*/\\rho_c$ and $\\delta = \\zeta_0/r$ so defining $\\epsilon= F_0 r /(\\rho_c \\Omega^2\\zeta_0^3).$ For a uniform response, one requires that $\\epsilon \\delta ^2 \\ll \\lambda.$ This is equivalent to the requirement that the product of the ratio of the external radiation momentum flux to the local disc pressure with the disc aspect ratio should be significantly less than unity. We recall here that because of the geometrical configuration, the momentum flux locally reradiated by the disc is in general much less than the external momentum flux at that location. In this paper we extend the treatment of disc warping induced by the action of an external pressure to include the effects of the response of the vertical structure particularly when this is non uniform as is the case when the above condition is not satisfied. In that case we find that the disc dynamics enters a different regime. This is such that the exposed surface tends to align so as to reduce the momentum absorption and hence the applied pressure. In the extreme limit of this regime the upper surface acts as if it is in contact with a rigid wall with the tendency to radiation warping instability tending to vanish. In this limit quantities characterising the disc twist and warp tend to oscillate at typical frequency $\\sim \\lambda \\Omega $. To illustrate these effects we derive a description of the one dimensional evolution of the disc inclination in radius and time that incorporates the effects of the vertical structure response, radiation torques and which applies both to the high viscosity regime, when the evolution is diffusive, and to the low viscosity regime when the evolution is wavelike. In all cases, the efficacy of surface radiation pressure driven instabilities is found to be reduced once the model parameters are such that the vertical response is significantly non uniform. We go on to estimate conditions for the response to be nonlinear and investigate the development of shock waves in the response using a one dimensional slab analogue model which has the same characteristic behaviour of linear perturbations as the full disc model. We find that such shocks potentially provide an important dissipation mechanism. The plan of this paper is as follows. In section \\ref{sec1} we describe some aspects of the thin disc model used. In section \\ref{Coords} we go on to introduce the twisted coordinate system used together with the notation convention, giving the basic equations in section \\ref {Beq}. By integrating over the vertical direction we use these to derive a single equation governing the dynamics of a twisted disc in section \\ref{Derveq}. When the disc behaves like a set of rigid rings for which the vertical response is uniform, this equation can be used to give a complete specification of the warp evolution. We note that this provides an extension of previous formulations to be able to consider the case when warps propagate as waves rather than diffuse radially (e.g. Nelson \\& Papaloizou 1999). In this Paper we use the twisted coordinate system formalism first introduced by Petterson 1977a, 1978 where the dynamical equations take the most simple form. When this formalism is adopted and an accurate description of all components of the equations of motion is needed as in the problem on hand we show that, in general, there is an ambiguity in choice of twisted coordinates with a set of these describing the same physical situation. As discussed in section \\ref{Derveq} a choice of a most appropriate twisted coordinate system can be motivated by the condition that perturbations of all dynamical quantities determined by the disc twist and warp are small. The transformation law between different twisted coordinates corresponding to the same physical situation is derived in appendix C for an inviscid disc. In order to obtain a complete description of the warp evolution when the vertical response is non uniform, we begin by obtaining a complete solution of the vertical problem for a polytropic model in section \\ref{poly}. This is used to obtain a pair of equations governing the radial warp evolution. We also indicate how the results can be extended to apply to more general models. We confirm the condition for the disc response to be like that of a series of rigid rings as $\\lambda \\gg \\epsilon (\\zeta_0/r)^2.$ We go on to perform a linear stability analysis of the radial evolution equations adopting a WKB approach in section \\ref {qualan} obtaining a maximum potential growth/decay rate in section \\ref{7.3}. In section \\ref{8} we discuss and confirm the analogy between the response of the disc and the linear and non linear dynamics of a vertically stratified one dimensional slab. We consider the development of shock waves and the formation of a rarefied hot atmosphere in section \\ref{8.3} giving a crude estimate of the warp dissipation rate in section \\ref{8.4}. Finally in section \\ref {Conc} we discuss and summarize our results. ", "conclusions": "\\label{Conc} In this paper we have presented a self-consistent calculation of the response of a twisted disc to the action of a radiation pressure force acting in upper layers of the disc. This is assumed to be due to the interception and subsequent reradiation of radiation from a central source. The radiation pressure force is assumed to be applied at a single density level $\\rho_{*}$ corresponding to optical thickness unity. The degree of twist and warping is assumed to be small enough that linear theory can be used to calculate the response. The analysis of Pringle 1996 modelled the disc as a set of concentric rigid rings in a state of Keplerian rotation. These were assumed to communicate with each other through the exchange angular momentum because of the action of viscous forces. Up to now there has been no consideration of effects arising from the response of the disc vertical structure to the externally applied pressure. In this paper we have considered the warping and twisting of an accretion disc taking account the response of the disc vertical structure to an external radiation field assuming that the effect of self-shielding of radiation by the global twisted disc is not significant. We developed a description of the evolution of the disc in terms of a pair of equations governing the one dimensional evolution of the inclination as a function of radius and time (see section \\ref{sec6} and equations (\\ref{e26}), (\\ref{e48}) and (\\ref{e49})). This description extends earlier ones, so that the influence of radiation pressure on discs for which warps occur in the low viscosity bending wave regime, as well as the higher viscosity diffusive regime, may be considered. We found several qualitatively different dynamical regimes that may be realised in astrophysical discs. These are related to whether the character of the response of the disc vertical structure causes significant departures from what would be obtained for a rigid body. \\subsection{ Conditions for a quasi-rigid response} We found that to avoid such departures, the momentum flux due to radiation from the central source $F_0,$ should be smaller than a critical value, $F_{*},$ given by $F_{*}=(\\lambda /\\delta )P_{c},$ where $\\delta = \\zeta_0/r $ is the disc aspect ratio, $P_{c}$ is the disc mid plane pressure and it has been assumed that the the ratio $(\\lambda /\\delta )$ is small (see sections \\ref{Dyneq} and \\ref{7.1.2}). Then, when $F_0 > F_{crit}\\approx 0.1(\\delta /\\alpha )P_{c},$ warping instability of the disc becomes a possibility ( eg. Pringle, 1996). Thus for radiation driven warping of a disc that behaves like an ensemble of rigid rings, we find two requirements, namely that $F_{*} > F_{crit},$ together with $F_0 < F _{*}.$ These conditions together imply that the ration of the surface to mid plane density should be sufficiently large and such that $\\lambda > \\lambda_{crit}\\approx 0.1\\delta^{2}/\\alpha.$ \\subsection{ Large external radiation momentum flux} In the opposite limit of large $F_0 > max(F_{*}, F_{crit}),$ consideration of the disc structure in the vertical direction is essential as the vertical displacement ceases to be uniform. The perturbed gas motion in the disc is mainly determined by the vertical component of velocity and the perturbed quantities tend to oscillate at a characteristic frequency $\\omega = (\\rho_c \\zeta_0 / 2\\Sigma)\\lambda \\Omega $, with $\\Omega$ being the local Keplerian angular velocity. In this situation, vertical motion tends to be suppressed by the external pressure field on the irradiated side of the disc while warping motions persist in the bulk of the disc and on the shielded side of the disc (see sections \\ref{Dyneq} and \\ref{smalllam}). The disc surface intercepting the radiation tends to align parallel to the radiation rays, thus decreasing the amount of intercepted radiation, while the inclination angles associated with the disc mid plane and the opposite free boundary of the disc oscillate, being approximately equal to each other. In this limit the upper surface of optical thickness one plays the role of an impenetrable wall. Thus the development of instability of the inclination angle due to radiation pressure effects (e.g. Pringle 1996) would be inhibited in this limit. In principle, the presence of warping motions in this regime would be associated with variability on a large characteristic time scale $T_{ch}\\sim \\lambda^{-1}T_{K}$, where $T_{K}$ and $\\lambda $ are some 'typical' values of the orbital period and the density ratio. \\subsection{Limits on the WKB growth rate} Following Pringle 1996 we have considered possible instabilities using a WKB approach. As this neglects potentially important global effects and boundary conditions results are not definitive, nonetheless they should give a good indication of when instability could be possible. We find that when $F_{0} < F_{*}$ the growth rate increases with $F_{0}$ while in the formal limit $F_{0} \\rightarrow \\infty$ it approaches zero. Thus in the linear theory there is an upper limit for the growth rate of the instability of the inclination angle which is always $\\le (\\rho_c \\zeta_0/ \\Sigma)\\lambda\\Omega $ in the absence of viscosity (see section \\ref{7.3}). When present, viscosity acts to damp any instability at a rate $0.1\\delta^2\\Omega /\\alpha,$ leading again to the condition $\\lambda > \\lambda_{crit}\\approx 0.1\\delta^{2}/\\alpha$ for radiation warping instability to be feasible. \\subsection{Conditions for non-linearity} Our results described above rely upon the applicability of linear perturbation theory. As discussed in section \\ref{7.5}, the breakdown of linear theory is expected when the Lagrangian change of pressure induced in the surface layers becomes of the order of the unperturbed pressure even though the change of inclination angle could be very small. The corresponding constraints on the inclination angle are especially strong for disc models with a significant drop of temperature toward the boundaries of the disc. \\subsection{Non linear calculations for the one dimensional slab analogue} A direct numerical approach to the issues discussed above is not yet feasible due to three-dimensional nature of the problem and the many physical processes involved. However, when vertical motions dominate, the problem becomes analogous to the one dimensional problem of vertical motions induced in a stratified gas column orbiting around a central source with Keplerian angular velocity. As discussed in Section \\ref{8} the one dimensional slab gives the same fundamental period of oscillation as obtained from the full disc theory when an appropriate boundary condition on the upper surface of the column is specified. The dependence of the eigenfrequency on theoretical parameters as well as the effective presence, in the limit of small $\\lambda,$ of an impenetrable wall at the upper surface of the column have been checked numerically. Where they can be compared, agreement between our analytical and numerical results is very good. We also used the one dimensional model in order to investigate possible consequences of non-linear behaviour in the system. To do this we studied the motion of the slab ensuing from the imposition of a vertical velocity profile with varying amplitude. We found that when the ratio of the Lagrangian pressure perturbation to the local value of the pressure at the upper boundary of the disc, $\\Delta P/P_{*}$, becomes larger than $1-10,$ strong shocks propagating downward into the column are observed (see section \\ref{8.3}). In principle, these shocks may lead to non-linear dissipation of energy stored in the vertical motion and in section \\ref{8.4} we made a very crude estimate of a possible time scale of $200$ orbits for $\\Delta P/P_{*}\\sim 10$. However, this result must be checked in framework of a more sophisticated numerical approach which takes into account physical processes neglected in this study, most notably the effects of radiation transfer in the upper layers of the column. \\subsection{Issues for future consideration} In a fully self-consistent study the dynamical effects of radiation pressure must be studied together with effects determined by the radiation heating of the disc atmosphere. This can be done in the most convenient way within the framework of the one dimensional model discussed above. In a realistic disc model, where effects due to the flaring of the disc lead to the interception and scattering of radiation in the disc photosphere, when axisymmetric and unperturbed, radiation heating can significantly influence on or even determine the value of the density ratio $\\lambda.$ This parameter, being the ratio of the density at the absorbing surface to the mid plane density defines the boundary between different dynamical regimes for a twisted disc. In this connection it is interesting to note that in certain vertical models of X-ray heated accretion discs, the ratio $\\lambda $ can be quite small, being of order of $10^{-4}-10^{-5}$, e.g. Jimenez-Garate et al 2002. In such models the new dynamical effects discussed in this Paper would certainly play an important role." }, "0710/0710.0686_arXiv.txt": { "abstract": "A survey of currently known extrasolar planets indicates that close to 20\\% of their hosting stars are members of binary systems. While the majority of these binaries are wide (i.e., with separations between 250 and 6500 AU), the detection of Jovian-type planets in the three binaries of $\\gamma$ Cephei (separation of 18.5 AU), GL 86 (separation of 21 AU), and HD 41004 (separation of 23 AU) have brought to the forefront questions on the formation of giant planets and the possibility of the existence of smaller bodies in moderately close binary star systems. This paper discusses the late stage of the formation of habitable planets in binary systems that host Jovian-type bodies, and reviews the effects of the binary companion on the formation of Earth-like planets in the system's habitable zone. The results of a large survey of the parameter-space of binary-planetary systems in search of regions where habitable planets can form and have long-term stable orbits are also presented. ", "introduction": " ", "conclusions": "" }, "0710/0710.2226_arXiv.txt": { "abstract": "Gamma Ray Bursts (GRBs) show evidence of different spectral shapes, light curves, duration, host galaxies and they explode within a wide redshift range. However, the most of them seems to follow very tight correlations among some observed quantities relating to their energetic. If true, these correlations have significant implications on burst physics, giving constraints on theoretical models. Moreover, several suggestions have been made to use these correlations in order to calibrate GRBs as standard candles and to constrain the cosmological parameters. \\\\ We investigate the cosmological relation between low energy $\\alpha$ index in GRBs prompt spectra and the redshift $z$. We present a statistical analysis of the relation between the total isotropic energy $E_{iso}$ and the peak energy $E_p$ (also known as Amati relation) in GRBs spectra searching for possible functional biases. \\\\ Possible implications on the $E_{iso}$ vs $E_p$ relation of the $\\alpha$ vs $(1+z)$ correlation are evaluated. We used MonteCarlo simulations and the boostrap method to evaluate how large are the effects of functional biases on the $E_{iso}$ vs $E_p$. We show that high values of the linear correlation coefficent, up to about 0.8, in the $E_{iso}$ vs $E_p$ relation are obtained for random generated samples of GRBs, confirming the relevance of functional biases. \\\\ Astrophysical consequences from $E_{iso}$ vs $E_p$ relation are then to be revised after a more accurate and possibly bias free analysis. ", "introduction": "Gamma-ray Bursts (GRBs) are brief and intense flashes of high energy radiation emitted mostly in the $\\gamma$-ray band. They are detected from wholly random directions in the sky at the rate of about once a day and typically last from a few milliseconds to several minutes. Within a few years, the BATSE experiment on board the NASA's Compton Gamma Ray Observatory satellite (Fishman et al. 1989) has recorded over 2700 GRB events with an isotropic distribution in the sky (Meegan et al. 1996). However, although BATSE was very sensitive to high-energy photons, it could not discern the direction of a burst to better than a few degrees uncertainty, too large to pinpoint the location of individual explosions. The real revolutionary step forward occurred in the 1997, thanks to the Italian-Dutch BeppoSAX satellite (Boella et al. 1997). This satellite was not as sensitive as BATSE to $\\gamma$ rays, but its relative quick response of pointing system, coupled with good accuracy position information, permitted the first detection of an X-ray {\\it afterglow}, the radiation emitted after the initial burst of $\\gamma$-ray (Costa et al. 1997). This discovery of afterglows made redshift measurements possible and confirmed that GRBs lie at cosmological distance ($0.0085$ (Galama et al. 1999) $ 200$ days) is similar to SN 2002cx, except for higher expansion velocities and higher velocity dispersions. The presence of P-Cygni profiles in the late phase spectrum of SN 2005hk indicates that the ejecta have not become optically thin till our last observation. Modeling of the pre-maximum spectra of SN 2005hk indicates a relatively higher temperature which makes \\ion{Fe}{iii} lines strong. The presence of weak \\ion{O}{i} line at $\\lambda$7774 at almost all epochs is modeled as a consequence of high abundance of completely mixed unburned oxygen in the ejecta. The late phase spectra of SN 2005hk are modeled as a combination of the photospheric and nebular components, with the nebular line emitting region being more centrally concentrated than is expected in the case of a lower energetic model." }, "0710/0710.1069_arXiv.txt": { "abstract": "We utilize existing imaging and spectroscopic data for the galaxy clusters MS2137-23 and Abell 383 to present improved measures of the distribution of dark and baryonic material in the clusters' central regions. Our method, based on the combination of gravitational lensing and dynamical data, is uniquely capable of separating the distribution of dark and baryonic components at scales below 100 kpc. Our mass models include pseudo-elliptical generalized NFW profiles for constraining the inner dark matter slope, and our lens modeling takes into account both the ellipticity and substructure associated with cluster galaxies as necessary in order to account for the detailed properties of multiply-imaged sources revealed in Hubble Space Telescope images. We find a variety of strong lensing models fit the available data, including some with dark matter profiles as steep as expected from recent simulations. However, when combined with stellar velocity dispersion data for the brightest member, shallower inner slopes than predicted by numerical simulations are preferred, in general agreement with our earlier work in these clusters. For Abell 383, the preferred shallow inner slopes are statistically a good fit only when the multiple image position uncertainties associated with our lens model are assumed to be 0\\farcs5, to account for unknown substructure. No statistically satisfactory fit was obtained matching both the multiple image lensing data and the velocity dispersion profile of the brightest cluster galaxy in MS2137-23. This suggests that the mass model we are using, which comprises a pseudo-elliptical generalized NFW profile and a brightest cluster galaxy component may inadequately represent the inner cluster regions. This may plausibly arise due to halo triaxiality or by the gravitational interaction of baryons and dark matter in cluster cores. The intriguing results for Abell 383 and MS2137-23 emphasize the need for a larger sample of clusters with radial arcs. However, the progress made via this detailed study highlights the key role that complementary observations of lensed features and stellar dynamics offer in understanding the interaction between dark and baryonic matter on non-linear scales in the central regions of clusters. ", "introduction": "Cold dark matter (CDM) simulations (both with and without the inclusion of baryonic physics) are a crucial tool and proving ground for understanding the physics of the universe on nonlinear scales. One of the most active aspects of research in this area concerns the form of the dark matter density profile. Key questions raised in recent years include: Is there a universal dark matter density profile that spans a wide range of halo masses? What is the form of this profile and how uniform is it from one halo to another? To what extent do baryons modify the dark matter distribution? Drawing on a suite of N-body simulations, \\citet{NFW97} originally proposed that the dark matter density profiles in halos ranging in size from those hosting dwarf galaxies to those with galaxy clusters have a universal form. This 3-D density distribution, termed the NFW profile, follows $\\rho_{DM}\\propto r^{-1}$ within some scale radius, $r_{sc}$, and falls off as $\\rho_{DM}\\propto r^{-3}$ beyond. Subsequent simulations indicated that the inner density profile could be yet steeper - $\\rho_{DM}\\propto r^{-1.5}$ \\citep{M98,Ghigna00}. As computing power increases and numerical techniques improve, it is now unclear whether the inner dark matter distribution converges to a power law form rather than becoming progressively shallower in slope at smaller radii \\citep{P03,Navarro04,Diemand04,Diemand05}. For comparisons with data, such simulations need to account for the presence of baryons. This is particularly the case in the cores of rich clusters. Although baryons represent only a small fraction of the overall cluster mass, they may be crucially important on scales comparable to the extent of typical brightest cluster galaxies. Much work is being done to understand the likely interactions between baryons and DM \\citep{Gnedin04,Nagai05,Faltenbacher05}. These simulations will provide refined predictions of the relative distributions of baryons and DM. This paper is a further step in a series which aims to present an observational analog to progress described above in the numerical simulations. At each stage it is desirable to confront numerical predictions with observations. Whereas some workers have made good progress in constraining the {\\em total} density profile (e.g.~\\citet{Broadhurst05b}), in order to address the relevance of the numerical simulations we consider it important to develop and test techniques capable of separating the distributions of dark and baryonic components (e.g. ~\\citet{Sand02,Zappacosta06,Biviano06,Mahdavi07}). This paper presents a refined version of the method first proposed by \\citet{Sand02}, exploited more fully in \\citet{Sand04} (hereafter S04). S04 sought to combine constraints from the velocity dispersion profile of a central brightest cluster galaxy (BCG) with a strong gravitational lensing analysis in six carefully selected galaxy clusters in order to separate the baryonic and dark matter distributions. S04 carefully selected clusters to have simple, apparently 'relaxed' gravitational potentials in order to match broadly the 'equilibrium' status of the simulated dark matter halos originally analyzed by \\citet{NFW97} and subsequent simulators. For example, Abell 383 and MS2137-23 have almost circular BCGs ($b/a$=0.90 and 0.83 respectively), require a single cluster dark matter halo to fit the strong lensing constraints (in contrast to the more typical clusters that require a multi-modal dark matter morphology -- Smith et al. 2005), have previously published lens models with a relatively round dark matter halo ($b/a$=0.88 and 0.78 respectively - Smith et al. 2001; Gavazzi 2005), and display no evidence for dynamical disturbance in the X-ray morphology of the clusters (Smith et al. 2005; Schmidt \\& Allen 2006). The merit of the approach resides in combining two techniques that collectively probe scales from the inner $\\sim$10 kpc (using the BCG kinematics) to the $\\sim$100 kpc scales typical of strong lensing. Whereas three of the clusters contained tangential arcs, constraining the total enclosed mass within the Einstein radius, three contained both radial and tangential gravitational arcs, the former providing additional constraints on the derivative of the total mass profile. In their analysis, S04 found the gradient of the inner dark matter density distribution varied considerably from cluster to cluster, with a mean value substantially flatter than that predicted in the numerical simulations. S04 adopted a number of assumptions in their analysis whose effect on the derived mass profiles were discussed at the time. The most important of these included ignoring cluster substructure and adopting spherically-symmetric mass distributions centered on the BCG. The simplifying assumptions were considered sources of systematic uncertainties, of order 0.2 on the inner slope. Although the six clusters studied by S04 were carefully chosen to be smooth and round, several workers attributed the discrepancy between the final results and those of the simulations as likely to arise from these simplifying assumptions \\citep{Bartelmann04,Dalal04b,Meneghetti05}. The goal of this paper is to refine the data analysis for two of the clusters (MS2137-23 and Abell 383) originally introduced by S04 using fully 2-D strong gravitational lensing models, thus avoiding any assumptions about substructure or spherical symmetry. The lensing models are based on an improved version of the LENSTOOL program (\\citealt{Kneibphd,Kneib96}; see Appendix; http://www.oamp.fr/cosmology/lenstool/). A major development is the implementation, in the code, of a pseudo-elliptical parameterization for the NFW mass profile, i.e. a generalization of those seen in CDM simulations, viz: \\begin{equation}\\label{eqn:gnfw} \\label{eq:gnfw} \\rho_d(r)=\\frac{\\rho_{c} \\delta_{c}}{(r/r_{sc})^{\\beta}(1+(r/r_{sc}))^{3-\\beta}} \\end{equation} where the asymptotic DM inner slope is $\\beta$. This formalism allows us to overcome an important limitation of previous work and takes into account the ellipticity of the DM halo and the presence of galaxy-scale subhalos. Furthermore the 2-D lensing model fully exploits the numerous multiply-imaged lensing constraints available for MS2137-23 and Abell 383. The combination of gravitational lensing and stellar dynamics is the most powerful way to separate baryons and dark matter in the inner regions of clusters. However, it is important to keep a few caveats in mind. Galaxy clusters are structurally heterogeneous objects that are possibly not well-represented by simple parameterized mass models. To gain a full picture of their mass distribution and the relative contribution of their major mass components will ultimately require a variety of measurements applied simultaneously across a range of radii. Steps in this direction are already being made with the combined use of strong and weak gravitational lensing (e.g.~\\citet{Limousin06,Bradac06}), which may be able to benefit further from information provided from X-ray analyses (e.g.~\\citet{Schmidt06}) and kinematic studies (e.g.~\\citet{Lokas03}). A recent analysis has synthesized weak-lensing, X-ray and Sunyaev-Zeldovich observations in the cluster Abell 478 -- similar cross-disciplinary work will lend further insights into the mass distribution of clusters \\citep{Mahdavi07}. Of equal importance are mass models with an appropriate amount of flexibility and sophistication. For instance, incorporating models that take into account the interaction of baryons and dark matter may shed light into the halo formation process and provide more accurate representations of dark matter structure. Halo triaxiality, multiple structures along the line of sight and other geometric effects will also be important to characterize. At the moment, incorporating these complexities and securing good parameter estimates is computationally expensive even with sophisticated techniques such as the Markov-Chain Monte Carlo method. Numerical simulation results are often presented as the average profile found in the suite of calculations performed. Instead, the distribution of inner slopes would be a more useful quantity for comparison with individual cluster observations. Also, comparisons between simulations and observations would be simplified if {\\it projected} density profiles of simulated halos along multiple lines of sight were to be made available. These issues should be resolvable as large samples of observed mass profiles are obtained. For the reasons above, comparing observational results with numerical simulations is nontrivial. The observational task should be regarded as one of developing mass modeling techniques of increasing sophistication that separate dark and baryonic matter, so as to provide the most stringent constraints to high resolution simulations which include baryons as they also increase in sophistication. The combination of stellar dynamics and strong lensing is the first crucial step in this process. Its diagnostic power will be further enhanced by including other major mass components (i.e.~the hot gas of the intracluster medium or the stellar contribution from galaxies) out to larger radii. A plan of the paper follows. In \\S~\\ref{sec:methods} we explain the methodology used to model the cluster density profile by combining strong lensing with the BCG velocity dispersion profile. In \\S~\\ref{sec:obsresults} we focus on translating observational measurements into strong lensing input parameters. This section includes the final strong lensing interpretation of MS2137-23 and Abell 383. In \\S~\\ref{sec:stronglens} we present the results of our combined lensing and dynamical analysis. In \\S~\\ref{sec:systematics} we discuss further systematic effects, limitations and degeneracies that our technique is susceptible to -- with an eye towards future refinements. Finally, in \\S~\\ref{sec:finale} we summarize and discuss our conclusions. Throughout this paper, we adopt $r$ as the radial coordinate in 3-D space and $R$ as the radial coordinate in 2-D projected space. When necessary, we assume $H_{0}$=65 \\kms Mpc$^{-1}$, $\\Omega_{m}$=0.3, and $\\Omega_{\\Lambda}$=0.7. ", "conclusions": "\\label{sec:systematics} In the previous section we have presented the results of our analysis, which showed that a mass model comprising a stellar component for the BCG following a Jaffe profile together with a generalized NFW DM cluster halo is able to adequately reproduce the observations for Abell 383 (albeit {\\it only} for the coarse lensing positional accuracy scenario) but is unable to simultaneously reproduce the observed multiple image configuration and BCG velocity dispersion profile for MS2137-23. In the case of Abell 383, the inner DM profile is flatter than $\\beta$=1, supporting the earlier work of S04. This indicates that at least some galaxy clusters have inner DM slopes which are shallower than those seen in numerical simulations -- but only if the mass parameterization used in the current work is reflective of reality. Further work in this interesting cluster using other observational probes will further refine the mass model, and determine if the generalized NFW DM form is a good fit to the cluster profile. In this section we discuss systematic uncertainties in our method and possible refinements that could be made to reconcile the mass model with the observations for MS2137-23. We hope that many of these suggestions will become important as cluster mass models improve and thus will present fruitful avenues of research. \\subsection{Systematic Errors} We focus first on systematic errors associated particularly with the troublesome stellar velocity dispersion profile for MS2137-23. Errors could conceivably arise from (i) significant non-Gaussianity in the absorption lines (which are fit by Gaussians), (ii) uncertain measurement of the instrumental resolution used to calibrate the velocity dispersion scale, and (iii) template mismatch. Non-gaussianity introduces an error that we consider too small to significantly alter the goodness of fit \\citep{Gavazzi05}. The instrumental resolution of ESI (the Keck II instrument used to measure the velocity dispersion profile; \\citet{Sheinis02}) is $\\sim$30 \\kms; this is much smaller than the measured dispersion. Even if the instrumental resolution was in error by a factor of two, the systematic shift in $\\sigma$ would only be 3 \\kms (using Eq~3 in Treu et al.\\ 1999). This would affect all measurements and not reverse the trend with radius. Concerning template mismatch, S04 estimated a possible systematic shift of up to 15-20 \\kms . This could play a role especially as the signal to noise diminishes at large radii, where the discrepancy with the model profile is greatest. To test this hypothesis, we added 20 \\kms in quadrature to only those velocity dispersion data points in MS2137-23 at $R > 4$ kpc and recalculated the best-fitting $\\chi^{2}$ values. $\\chi^{2}$ is reduced from 31 to 28.8, a modest reduction which fails to explain the poor fit. Although selectively increasing the error bars on those data points most discrepant with the model is somewhat contrived, our result does highlight the need for high S/N velocity dispersion measures out to large radii. A high quality velocity dispersion profile has been measured locally for Abell 2199 to $\\sim$20 kpc \\citep{Kelsonetal02}. Interestingly, these high S/N measures display similar trends to those for MS2137-23 in the overlap regime, i.e. a slightly decreasing profile at $R\\lesssim 10$kpc. The dip witnessed in MS2137-23 is thus not a unique feature, although with deeper measurements we might expect to see a rise at larger radii as a result of the shallow DM profile. A final potential limitation in the dynamical analysis is the assumption of orbital isotropy. Both S04 and \\citet{Gavazzi05} explored the consequences of mild orbital anisotropy, concluding a possible offset of $\\Delta \\beta \\sim0.15$ might result. Even including orbital anisotropy into his analysis, \\citet{Gavazzi05} was unable to fit the observed velocity dispersion profile. Since we determine our lensing $\\chi^{2}$ values in the source plane, we checked to make sure that no extra images were seen after remapping our best-fit lensing + velocity dispersion models back to the image plane. No unexpected images were found, although several images that were explicitly not used as constraints were found, such as the mirror image of radial arc image 2a in MS2137 and the complex of multiple images associated with 3abc, 5ab, and 6ab in Abell 383 (see Figures~\\ref{fig:mulplot} and \\ref{fig:mulplota383}). As discussed in \\S~\\ref{sec:lensinterpms2137} and \\ref{sec:lensinterpa383}, some of these multiple images were not used as constraints because we could not confidently identify their position either due to galaxy subtraction residuals or blending with other possible multiple image systems. \\begin{figure*} \\begin{center} \\mbox{ \\mbox{\\epsfysize=4.5cm \\epsfbox{f5a.eps}} \\mbox{\\epsfysize=4.5cm \\epsfbox{f5b.eps}} } \\mbox{ \\mbox{\\epsfysize=4.5cm \\epsfbox{f5c.eps}} \\mbox{\\epsfysize=4.5cm \\epsfbox{f5d.eps}} } \\caption{Confidence contours (68\\%,95\\%, and 99\\%) when we allow the dark matter scale radius to be fixed at values a factor of two beyond our observationally motivated prior. Top Row -- Contours when we fix the dark matter scale radius to $r_{sc}$=50 and 400 kpc in MS2137. Although the $r_{sc}$=400 kpc scenario provides a relatively good fit to the data ($\\chi^{2}\\sim$26), this value for the scale radius is much larger than that observed from weak lensing data. The $r_{sc}$=50 kpc scenario is a significantly worse fit to the data, with $\\chi^{2}\\sim$39. Note that the DM inner slope is $\\beta < 1$ in both scenarios. Bottom Row -- Contours when we fix the dark matter scale radius to $r_{sc}$=50 and 400 kpc in A383. The large discrepancy in inner slope values obtained emphasize the need for a mass probe at larger radii. The best-fitting model for either fixed scale radius is significantly worse than the best-fitting $r_{sc}$=100 kpc result ($\\chi^{2}\\sim$26.5 and 31.3 for $r_{sc}$=50 and 400 kpc respectively). \\label{fig:diffrsc}} \\end{center} \\end{figure*} We finally comment on the uncertainties assigned to the multiple image systems for our lens models. We have presented two sets of results in this work; with assigned image positional accuracies of $\\sigma_{I}$=0\\farcs2 and 0\\farcs5. We find a variety of lens models are compatible with the $\\sigma_{I}$=0\\farcs2 case and only when the velocity dispersion data is included into the analysis does the data fail to be reproduced by the model. Certainly if we were to further increase the positional errors, at some point a good velocity dispersion fit could conceivably be obtained, but we will refrain from doing so in the present work. Increasing the positional uncertainties is only justified if there is evidence that there are significant missing components in the mass models. Further observations that can probe the mass distribution on fine scales to larger radii and higher quality models which can account for phenomena such as adiabatic contraction in the inner regions of galaxy clusters and triaxiality represent the best way to obtain a more precise picture of the cluster mass distribution. \\subsection{Improving the Mass Model} We now turn our attention to possible inadequacies in the mass model. It is important to stress that the two diagnostics (lensing and dynamics) adopted in this study probe different scales. The lensing data tightly constrains the mass profile at and outside the radial arc ($\\sim$20 kpc), while the velocity dispersion constrains the mass profile inside $R \\lesssim 10$ kpc. Since multiple images are numerous and their positions can be more precisely measured than velocity dispersion \\footnote{The error on the astrometry with respect to the relevant scale, the Einstein Radius $\\theta_{\\rm E}$ is much smaller than the relative error on velocity dispersion, i.e. $\\delta \\theta / \\theta_{\\rm E} << \\delta \\sigma/\\sigma$}, they carry more weight in the $\\chi^2$ statistic than the kinematic points, producing a best overall fitting model (which is lensing dominated) that is a relatively poor fit to the kinematic data. To improve the model, one must admit that either one of the two components of the modeling is incorrect, or that the functional form of the mass profile chosen to extrapolate the lensing information at the scales relevant for dynamics is insufficient. In this section we discuss several areas where the mass model presented in this paper could be improved. \\subsubsection{The Contribution of the Brightest Cluster Galaxy} We might query the assumption of a Jaffe density profile for the BCG. This seems an unlikely avenue for improvement given the Jaffe profile fits the observed BCG surface brightness profile remarkably well (see Figure 2 of S04). Moreover, \\citet{Gavazzi05} utilized a Hernquist mass profile in his analysis of MS2137-23, which also matches the observations, and \\citet{Gavazzi05} was likewise unable to reproduce the observed S04 velocity dispersion profile. We have additionally checked our assumptions by altering the PIEMD fit to the BCG surface brightness data so that it is matched not to the derived Jaffe profile fit to the BCG but directly to the HST surface brightness profile. With this setup, we found a $r_{cut}$ value of 23.70 kpc for MS2137 and 28.65 kpc for Abell 383 (compare this with the numbers in Table~\\ref{tab:lensfixed}). Redoing our analysis for the best-fitting $r_{sc}$ scenario only, our constraints on $\\beta$ for both Abell 383 and MS2137 did not change by more than 0.05, and so it is not likely that our method for constraining the BCG mass contribution is the root cause of our inability to fit the data to a BCG + gNFW cluster DM halo mass model. Conceivably the BCG may not be coincident with the center of the cluster DM halo, as has been assumed throughout this work. It is often the case that small subarcsecond off sets between BCGs and cluster DM halos are necessary to fit lensing constraints (e.g.~\\citet{gps05}). There is strong evidence that the BCG is nearly coincident with the general cluster DM halo {\\it in projection} from the strong lensing work presented here and by others \\citep{gavazzi03,Gavazzi05}. However, an offset could be responsible for the flat to falling observed velocity dispersion profile if the BCG were actually in a less dark matter dominated portion of the cluster. Another possibility is that there are multiple massive structures along the line of sight, which would be probed by the strong lensing analysis, but not with the velocity dispersion profile of the BCG. A comprehensive redshift survey of MS2137-23 could provide further information on structures along the line of sight. \\subsubsection{The Advantage of a Mass Probe at Larger Radii}\\label{sec:highrad} With our presented data set, we have seen that it is difficult to constrain the DM scale radius, $r_{sc}$ because both of our mass probes are only effective within the central $\\sim$100 kpc of the clusters -- within the typical DM scale radius observed and seen in CDM simulations. For this reason, the inferred DM scale radius for both Abell 383 and MS2137-23 lay at the boundary of our assumed prior range. Future work will benefit from weak lensing data, along with galaxy kinematics and X-ray data of the hot ICM which can each probe out to large clustercentric radii. Although not the focus of the current work, pinning down the correct DM scale radius will be crucial for constraining other DM mass parameters. For instance, there is a well-known degeneracy between $r_{sc}$ and the inner slope $\\beta$ (e.g.~\\citet{gavazzi03,Gavazzi05}). To briefly explore this, we have reran our analysis (for the coarse positioning lensing case) for both clusters with a $r_{sc}$ of 50 and 400 kpc -- factors of two beyond our chosen $r_{sc}$ prior. We show our confidence contours in Figure~\\ref{fig:diffrsc}, which are noteworthy. For example in the case of MS2137-23, if we fix $r_{sc}$=50 kpc, then the best-fitting $\\beta = 0.05$. However, if $r_{sc}$=400 kpc then $\\beta=0.7$, more in accordance with simulations. Interestingly, the $r_{sc}$=400kpc scenario returns a better overall $\\chi^{2}\\sim26$ than any model with $r_{sc}$=100-200 kpc -- even though a $r_{sc}$ of 400 kpc is clearly ruled our by extant weak lensing observations. None of the other $r_{sc}$=50,400 kpc scenarios produced $\\chi^{2}$ values that were comparable to those seen with $r_{sc}$=100-200 kpc. Any further knowledge of the DM scale radius would aid greatly in constraining $\\beta$ and determining the overall goodness of fit of the generalized NFW DM profile to the cluster data. X-ray studies assuming hydrostatic equilibrium \\citep{Allen01vir,Schmidt06} and a combined strong and weak lensing analysis \\citep{Gavazzi05} have presented data on MS2137-23 to radii much larger than that probed in this study. To check that the mass model derived from data within $\\sim$ 100 kpc do not lead to results at variance with published data at larger radii, we have taken the \\citet{Gavazzi05} results and compared their derived mass at large radii with an extrapolation of our mass models. Examining Figure~3 from \\citet{Gavazzi05} we estimate that from his weak lensing analysis a 2D projected mass enclosed between $1.6 \\times 10^{14}$ and $1.1 \\times 10^{15} M_{\\odot}$ at $\\sim 1.08$ Mpc using the cosmology adopted in this paper. Correspondingly, if we take all of the $\\Delta \\chi^{2}<1.0$ models using our analysis method (the coarse positional accuracy case was use) and calculate the expected 2D projected mass enclosed at 1.08 Mpc we find values between $6.9 \\times 10^{14}$ and $8.4 \\times 10^{14} M_{\\odot}$, well within the expected range. Note that no attempt was made to extrapolate the mass {\\it profiles} derived in our analysis to larger radii than the data in this paper allow, although we are acquring weak lensing data for a large sample of galaxy clusters to perform a more extensive analysis. The purpose of this consistency check is to only ensure that the masses we derive for such large radii are not too discrepant with existing analyses. The consistency check is satisfied and lends some credence to the models. \\subsubsection{Dark matter baryons interactions and Triaxiality}\\label{sec:ac} The central regions of DM halos can be strongly affected by the gravitational interaction with baryons during halo formation. If stars form and condense much earlier than the DM, it is expected that the baryons will adiabatically compress the DM resulting in a halo that is {\\it steeper} than that of the original \\citep{Blumenthal86,Gnedin04}. Alternatively, dark matter heating through dynamical friction with cluster galaxies can counteract adiabatic contraction, leading to a shallower DM profile \\citep{Elzant04,Nipoti03}. The present work takes into account neither of the above scenarios, and if any baryon-DM interaction greatly changes the cluster density profile, our assumed parameterized gNFW profile may be inappropriate. Recently, \\citet{Zappacosta06} have used X-ray mass measurements in the cluster Abell 2589 to conclude that processes in galaxy cluster formation serve to counteract adiabatic contraction in the cluster environment. Certainly, more observational work is needed to understand the interplay between baryons and DM in clusters, and extended velocity dispersion profiles of BCGs in conjunction with other mass tracers at larger radii could serve as the best testing ground for the interplay of dark and luminous matter. Not only is there likely significant interplay between baryons and DM in the central regions of clusters, but real galaxy clusters are certainly triaxial and, if ignored, this may lead to biased parameter estimations and discrepancies when combining mass measurement techniques that are a combination of two- and three-dimensional. Several recent studies have considered the effects of halo triaxiality on observations. Using an N-body hydrodynamical simulation of a disk galaxy and performing a 'long slit' rotation curve observation, \\citet{Hayashi04} found that orientation and triaxial effects can mistake a cuspy DM profile for one that has a constant density core. At the galaxy cluster scale, \\citet{Clowe04} performed mock weak lensing observations of simulated galaxy clusters and found that the NFW concentration parameter recovered was correlated with the 3D galaxy cluster orientation. In order to investigate the recent rash of galaxy clusters with observed high concentration parameters in seeming contradiction to the CDM paradigm \\citep{Kneib03,Gavazzi05,Broadhurst05b}, \\citet{Oguri05} used strong and weak lensing data in Abell 1689 along with a set of models that included halo triaxiality and projection effects. Again, it was seen that halo shape causes a bias in mass (and mass profile) determination, although it should be kept in mind that measurements of concentration are extremely difficult (e.g. Halkola et al.\\ 2006), and the recent study of \\citet{Limousin06} has seemed to clear up the concentration parameter controversy for at least Abell 1689. In terms of the current work, \\citet{Gavazzi05} has pointed out that the inability of his lensing model to fit the MS2137-23 BCG velocity dispersion profile may be due to halo triaxiality or projected mass along the line of sight (which would increase the mass measured in the lensing analysis but would not be seen in the stellar velocity dispersion). \\citet{Gavazzi05} showed that an idealized prolate halo with an axis ratio of $\\sim$ 0.4 could explain the velocity dispersion profile in MS2137-23. Halo triaxiality could also explain the high concentration previously seen in this cluster. Again, the gap between simulations and observations may be bridged with respect to triaxiality if further steps were taken to compare the two directly. One step in this direction would be the publication of detailed density profiles for the simulations (in 3-D or along numerous projected sight-lines). The most recent DM only simulations have indicated that the standard NFW profile representation of a DM profile (and its \\citet{M99} counterpart with an inner slope $\\beta \\sim 1.5$) can be significantly improved by slightly altering the model to a profile with a slope that becomes systematically shallower at small radii (e.g.~\\citet{Navarro04}, but see \\citet{Diemand05}). While we have adopted the traditional generalized NFW profile in this study, future work with parameterized models should move towards the latest fitting functions along with an implementation of adiabatic contraction as has already been attempted by \\citet{Zappacosta06}. Note, however, that both \\citet{Navarro04} and \\citet{Diemand04} have stated that all fitted functions have their weaknesses when describing complicated N-body simulations and that when possible simulations and observations should be compared directly." }, "0710/0710.1891_arXiv.txt": { "abstract": "Our high time resolution observations of individual pulses from the Crab pulsar show that the main pulse and interpulse differ in temporal behavior, spectral behavior, polarization and dispersion. The main pulse properties are consistent with one current model of pulsar radio emission, namely, soliton collapse in strong plasma turbulence. The high-frequency interpulse is quite another story. Its dynamic spectrum cannot easily be explained by any current emission model; its excess dispersion must come from propagation through the star's magnetosphere. We suspect the high-frequency interpulse does not follow the ``standard model'', but rather comes from some unexpected region within the star's magnetosphere. Similar observations of other pulsars will reveal whether the radio emission mechanisms operating in the Crab pulsar are unique to that star, or can be identified in the general population. ", "introduction": "Between 1994 and 2002 our group observed strong pulses from the Crab pulsar between 1 and 5 GHz at the VLA and Arecibo. Our Arecibo observations were designed with 2-ns time resolution, in order to test competing models of the radio emission mechanism. We initially concentrated on the MP, because it is stronger than the IP below $\\sim$5 GHz in the star's mean profile, and also because giant pulses are more common at the phase of the MP at these frequencies \\cite{Cordes2004}. Our results, and our data acquisition system, are described in \\cite{HKWE}. \\begin{figure}[ht] \\vspace{-1.5in} \\rotatebox{-90}{ \\includegraphics[trim = 40 0 60 0, width=0.6\\textwidth,clip]{EilekFig1.ps}} \\vspace{-2.5in} \\caption{An example of a Main Pulse, observed with 2.2-GHz bandwidth at 9 GHz, and coherently dedispersed \\cite{HKWE}. The pulse seen in total intensity (upper panel, plotted with total intensity time resolution 6.4 ns) contains several short-lived microbursts. The dynamic spectrum (lower panel, plotted with resolution 102 ns and 19.5 MHz) shows that the microburst emission spans the full receiver bandwidth. In a few MPs individual, short-lived nanoshots are sparse enough in time to be separately identified (shown in \\cite{HKWE, HE2007}); these examples reveal the dynamic spectrum of the nanoshots is relatively narrow. } \\label{MPfig} \\end{figure} \\begin{figure}[ht] \\rotatebox{-90}{ \\includegraphics[trim = 40 0 60 0, width=0.6\\textwidth,clip]{EilekFig2.ps}} \\caption{An example of an Interpulse, observed with 2.2 GHz bandwidth at 9 GHz, and coherently dedispersed. The IP seen in total intensity (upper panel) typically contains 1 or 2 sub-bursts; thus it has a simpler time signature than the MP (as in the example of Figure \\ref{MPfig}). The regular {\\it emission bands} in the dynamic spectrum (lower panel) are not due to instrumental or interstellar effects, but are characteristic to the emission physics of the IP. Note that the secondary burst, seen in total intensity, coincides with the appearance of new band sets in the dynamic spectrum. Plotted with total intensity time resolution 51.2 ns, and dynamic spectrum resolution 104 ns and 19.5 MHz. } \\label{IPfig} \\end{figure} To follow up on these results, we extended our data acquisition system to higher time resolution. We went to higher frequencies (5 to 10 GHz) in order to take advantage of the 2.2-GHz bandwidths (and corresponding sub-ns time resolution) available at Arecibo. We recorded individual IPs as well as MPs, because at these frequencies strong pulses are much more common at the phase of the IP \\cite{Cordes2004}. These new observations, carried out between 2003 and 2006, are reported in \\cite{HE2007}. We were astonished to find that IPs are very different from MPs at these frequencies. The IP differs from the MP in polarization, time signature, spectrum and dispersion, as discussed below. All of our IP observations were taken betweeen 5 and 10 GHz. Technical limitations, as well as the scarcity of strong IPs at lower frequencies, kept us from observing the IP below 4 GHz. We are therefore describing the ``high-frequency IP'', which occurs at a slighly earlier rotation phase from the ``regular'' IP (as seen at lower radio frequencies as well as in high-energy bands; {\\em e.g.}, \\cite{MH1996}). This phase offset suggests that the high-frequency IP may not be related at all to the regular IP; it may come from a very different part of the magnetosphere. ", "conclusions": "We have identified two different types of coherent radio emission from the Crab pulsar, one associated with the MP, the other associated with the high-frequency IP. Two different radiation mechanisms seem to be be operating within the star's magnetosphere. In addition, the higher dispersion of the high-frequency IP suggests the signal has passed through an unusually large plasma column before leaving the pulsar. Conventional wisdom ascribes the MP to emission from the open field line region above one of the star's magnetic axes. If this is the case, the high-frequency IP is probably not simply radiation from the other magnetic pole. It is more likely to come from some unexpected part of the magnetosphere; its phase offset, relative to the regular IP \\cite{MH1996}, corroborates this idea. The suggestion from \\cite{Maxim} that the high-frequency IP is emitted within the closed field line region may be on the right track." }, "0710/0710.2776_arXiv.txt": { "abstract": "{} {The nature of \\sori~(S\\,Ori~J053810.1$-$023626), a faint mid-T type object found towards the direction of the young \\so~cluster, is still under debate. We intend to disentangle whether it is a field brown dwarf or a 3-Myr old planetary-mass member of the cluster.} {We report on near-infrared $JHK_s$ and mid-infrared [3.6] and [4.5] IRAC/Spitzer photometry recently obtained for \\sori. The new near-infrared images (taken 3.82\\,yr after the discovery data) have allowed us to derive the first proper motion measurement for this object. } {The colors $(H-K_s)$, $(J-K_s)$ and $K_s$\\,$-$\\,[3.6] appear discrepant when compared to T4--T7 dwarfs in the field. This behavior could be ascribed to a low-gravity atmosphere or alternatively to an atmosphere with a metallicity significantly different than solar. The small proper motion of \\sori~(11.0\\,$\\pm$\\,5.9~mas\\,yr$^{-1}$) indicates that this object is further away than expected if it were a single field T dwarf lying in the foreground of the \\so~cluster. Our measurement is consistent with the proper motion of the cluster within 1.5\\,$\\sigma$ the astrometric uncertainty. } {Taking into account \\sori's proper motion and the new near- and mid-infrared colors, a low-gravity atmosphere remains as the most likely explanation to account for our observations. This supports \\sori's membership in \\so, with an estimated mass in the interval 2--7~\\mj, in agreement with our previous derivation.} ", "introduction": "Knowledge of the initial mass function is crucial for understanding the formation processes of stars, brown dwarfs and free-floating planetary-mass objects. Whether and where there is a limit for the creation of objects by direct collapse and fragmentation of molecular clouds has become one of the major goals in the study of very young populations. Planetary-mass candidates with masses in the interval 3--13 Jovian masses (\\mj) have been found in various star-forming regions (e.g., Lucas \\& Roche \\cite{lucas00}; Zapatero Osorio et al. \\cite{osorio00}; Chauvin et al. \\cite{chauvin04}; Lucas et al. \\cite{lucas05}; Luhman et al. \\cite{luhman05}; Jayawardhana \\& Ivanov \\cite{ray06}; Allers et al. \\cite{allers06}; Gonz\\'alez-Garc\\'\\i a et al. \\cite{gonzalez06}; Caballero et al. \\cite{caballero07}). These objects are mostly free-floating but in in a few cases appear as wide companions to young brown dwarfs or low-mass stars. \\sori~(S\\,Ori~J053810.1$-$023626) is the coolest free-floating, planetary-mass candidate so far reported in the literature. It was discovered by Zapatero Osorio et al$.$ (\\cite{osorio02a}) and lies in the direction of the \\so~cluster (352~pc and 1--8~Myr, with a best estimate at 3~Myr; Perryman et al. \\cite{perryman97}; Oliveira et al. \\cite{oliveira02}; Zapatero Osorio et al. \\cite{osorio02b}; Sherry et al. \\cite{sherry04}). The spectral type of \\sori~was determined at T5.5\\,$\\pm$\\,1.0 from molecular indices measured over near-infrared $H$- and $K$-band low-resolution spectra. Mart\\'\\i n \\& Zapatero Osorio (\\cite{martin03}) obtained an intermediate-resolution spectrum from 1.17 to 1.37~$\\mu$m ($J$-band), in which the K\\,{\\sc i} doublet at 1.25\\,$\\mu$m was detected. After comparison with theoretical spectra from Allard et al. (\\cite{allard01}), the authors inferred an effective temperature and surface gravity of $T_{\\rm eff}$\\,=\\,1100\\,$^{+200}_{-100}$~K and log\\,$g$\\,=\\,3.5\\,$\\pm$\\,0.5~cm~s$^{-2}$, in agreement with the expectations for a few megayears-old T dwarf. State-of-the-art evolutionary models (Chabrier \\& Baraffe \\cite{chabrier00}; Burrows et al. \\cite{burrows97}; Baraffe et al. \\cite{baraffe98}) yield a mass of 3\\,$^{+5}_{-1}$~\\mj~if \\sori's very young age is finally confirmed. Burgasser et al. (\\cite{burgasser04}), in contrast, have raised doubts about the low-gravity atmosphere and true cluster membership of \\sori. Based on the supposed similarity of the observed spectra to field T6--T7 dwarfs, these authors argued that the S\\,Ori object is ``an old, massive field brown dwarf lying in the foreground of the \\so~cluster''. However, this work relied on low signal-to-noise ratio data. Better quality photometry and spectra are needed to assess the true nature of this candidate. Here we present astrometric measurements, IRAC/Spitzer data and $JHK_s$ photometry for \\sori. We find that this object has colors unexpected for its spectral classification, which is measured in the range T4.5--T7 with a best estimate at T6. We ascribe this to a low gravity atmosphere, with a different metallicity being an alternative, but less likely, explanation. \\begin{table*} \\caption[]{Log of near-infrared observations of \\sori.} \\label{obslog} \\centering \\begin{tabular}{llcccc} \\hline\\hline Telescope & Instrument & Field of view & Pixel & Observing dates & Exposure time \\\\ & & (arcmin$^2$) & (arcsec) & & (s) \\\\ \\hline 3.5\\,m CAHA & Omega-2000 & 225 & 0.45 & 2005 Oct 22 ($JH$), 2005 Oct 25 ($K_s$) & 3600 ($J$), 6000 ($H$), 7200 ($K_s$) \\\\ 10\\,m KeckII& NIRSPEC & 0.59 & 0.18 & 2005 Oct 26 ($JK'$) & 405 ($J$), 270 ($K'$) \\\\ \\hline \\end{tabular} \\end{table*} \\begin{table*} \\caption[]{Near-infrared (2MASS photometric system) and IRAC/Spitzer photometry of \\sori.} \\label{phot} \\centering \\begin{tabular}{lccccccc} \\hline\\hline Telescope & $J$ & $H$ & $K_s$ & [3.6] & [4.5] & $H-K_s$ & $J-K_s$ \\\\ & (mag) & (mag) & (mag) & (mag) & (mag) & (mag) & (mag) \\\\ \\hline 3.5 m CAHA & 19.98\\,$\\pm$\\,0.06 & 20.07\\,$\\pm$\\,0.07 & 19.60\\,$\\pm$\\,0.08 & ... & ... & $+$0.48\\,$\\pm$\\,0.11 & $+$0.38\\,$\\pm$\\,0.10 \\\\ Keck\\,II & 19.96\\,$\\pm$\\,0.07 & ... & 19.58\\,$\\pm$\\,0.07 & ... & ... & ... & $+$0.38\\,$\\pm$\\,0.10 \\\\ Spitzer & ... & ... & ... & 18.62\\,$\\pm$\\,0.30 & 17.19\\,$\\pm$\\,0.15 & ... & ... \\\\ \\hline \\end{tabular} \\end{table*} ", "conclusions": "The low proper motion of \\sori~($\\mu$\\,=\\,11.0\\,$\\pm$\\,5.9~mas~yr$^{-1}$) makes it unlikely that it is a nearby ($\\le$30~pc) T dwarf. On the one hand, we compared our measurement with the motion of 192 Hipparcos stars (Perryman et al. \\cite{perryman97}) within a radius of $15^\\circ$ around \\so~and at a distance between 80 and 130~pc, which is the distance interval expected for \\sori~if it were a field, single T6 dwarf. About 70\\%~of the Hipparcos stars show larger motion than \\sori, suggesting that the S\\,Ori object is located farther away. On the other hand, our measurement is consistent (within 1.5~$\\sigma$) with the proper motion of the O9.5V--B0.5V-type star \\so~AB, which is the most massive member of the cluster of the same name. We note, however, that the relative motion of the Orion OB association is directed away from the Sun (de Zeeuw et al. \\cite{zeeuw99}). This makes it very hard to detect cluster members via proper motion analysis. On the contrary, radial velocity studies can be more discriminant (Jeffries et al. \\cite{jeffries06}) but the extreme faintness of \\sori~prevents any accurate radial velocity measurement with current instrumentation. \\subsection{Color-color diagrams} Color-color diagrams are depicted in Fig.~\\ref{colcol}. To put \\sori~into context, we included $JHK_s$ data of more than 100 field T-dwarfs and IRAC/Spitzer photometry of 36 field T-dwarfs compiled from the literature (Knapp et al. \\cite{knapp04}; Tinney et al. \\cite{tinney05}; Patten et al. \\cite{patten06}; Burgasser et al. \\cite{burgasser06a}; Artigau et al. \\cite{artigau06}; Mugrauer et al. \\cite{mugrauer06}; Leggett et al. \\cite{leggett99}, \\cite{leggett02}, \\cite{leggett07}; Liu et al. \\cite{liu07}; Luhman et al. \\cite{luhman07}; Looper et al. \\cite{looper07}). All near-infrared colors were conveniently transformed into the 2MASS photometric system using equations quoted in Stephens \\& Leggett (\\cite{stephens04}), which are appropriate for ultracool dwarfs. The location of \\sori~in Fig.~\\ref{colcol} is challenging since this object appears as an outlier, particularly when the $K$-band magnitude is involved. Two field dwarfs lie near it in the near-infrared color-color panels: 2MASS\\,J00501994$-$3322402 (Tinney et al. \\cite{tinney05}), whose photometric errors are quite large, and 2MASS\\,J13243559$+$6358284 (Looper et al. \\cite{looper07}). The latter object is widely discussed by Looper et al. (\\cite{looper07}) in terms of binarity and low-gravity atmosphere. We applied the criterion defined by Covey et al. (\\cite{covey07}, eq.~2) to distinguish objects with photometric properties deviating from the properties typical of field dwarfs and found that \\sori~lies more than 2\\,$\\sigma$ away from the near-infrared sequence defined by the field T-type brown dwarfs. In the mid-infrared wavelengths, the photometry of \\sori~deviates from the field on a 1--2\\,$\\sigma$ level. Covey et al.'s equation takes into account the color uncertainties of \\sori~and the width of the field distribution. As compared to T4--T7 field dwarfs, \\sori~presents redder $(H-K_s)$, $(J-K_s)$, and [3.6]\\,$-$\\,[4.5] colors and a bluer $K_s$\\,$-$\\,[3.6] index than expected for its spectral type ($\\sim$T6). However, the $(J-H)$ index is similar to that of T5--T6 field dwarfs. The $K$-band reddish nature of \\sori~is also apparent in its low-resolution $HK$ spectrum. Fig.~4 of Burgasser et al. (\\cite{burgasser04}) shows these data along with the spectrum of the field T6.5 2MASS\\,J10475385$+$2124234. Both spectra were obtained with similar instrumentation and are normalized to unity at 1.57~$\\mu$m. While the field dwarf matches reasonably well the $H$-band region of \\sori, it underestimates the flux at $K$-band, supporting the redder $(H-K_s)$ index of \\sori. Burgasser et al. (\\cite{burgasser04}) argued that this may be indicative of a ``lower surface gravity for \\sori~relative to 2MASS\\,J10475385$+$2124234''. Multiplicity cannot explain the observed photometric properties of \\sori. We artificially produced near-infrared colors of L--T and T--T pairs using the absolute magnitudes provided by Liu et al. (\\cite{liu06}). None of the combinations were able to reproduce our observations. The comparison of our data to theory is shown in the right panel of Fig.~\\ref{colcol}. The models depicted are those of Tsuji et al. (\\cite{tsuji04}), but in our analysis we also employed cloudless models by Marley et al. (\\cite{marley02}) and Burrows et al. (\\cite{burrows06}) obtaining similar results. The agreement between the models and the field dwarf observations is reasonably good. The great majority of the mid- and late-T dwarfs lie within the log\\,$g$\\,=\\,4.0 and 5.5~dex tracks, as expected for ``old'' dwarfs in the solar neighborhood. This is also consistent with the recent results of the spectral fitting work by Burgasser et al. (\\cite{burgasser06b}). These authors employed models by Burrows et al. (\\cite{burrows06}). The near-infrared photometry of \\sori~and current state-of-the-art theory of ultracool dwarfs indicate that this object may have a lower-gravity atmosphere than similarly classified T dwarfs in the solar vicinity. Because of the different pressure and density conditions at which H$_2$, CH$_4$, and CO absorptions are produced, low-gravity objects tend to be brighter at $K$ and redder in all near-infrared and [3.6]\\,$-$\\,[4.5] colors than comparable high-gravity objects (see discussions and Figures in Knapp et al. \\cite{knapp04}; Patten et al. \\cite{patten06}; Leggett et al. \\cite{leggett07}; Liebert \\& Burgasser \\cite{liebert07}). This is what we qualitatively observe in \\sori. From the right panel of Fig.~\\ref{colcol} and using the solar-metallicity models by Tsuji et al. (\\cite{tsuji04}), we derive log\\,$g$\\,$\\sim$\\,3.0 dex and $T_{\\rm eff}$\\,$\\sim$\\,1000--1100~K. This is consistent with previous results obtained from the spectral fitting analysis of low- and intermediate-resolution near-infrared spectra: $T_{\\rm eff}$\\,$\\sim$\\,800$^{+200}_{-100}$ K and log\\,$g$\\,$\\sim$\\,4.0\\,$\\pm$\\,1.0 dex (Zapatero Osorio et al. \\cite{osorio02a}), $T_{\\rm eff}$\\,$\\sim$\\,1100$^{+200}_{-100}$ K and log\\,$g$\\,$\\sim$\\,3.5\\,$\\pm$0.5 dex (Mart\\'\\i n \\& Zapatero Osorio \\cite{martin03}), respectively. These authors compared observations to theoretical data computed by Allard et al. (\\cite{allard01}). Recently, Liu et al. (\\cite{liu07}) have quantified the sensitivity of near-infrared spectra with $T_{\\rm eff}$\\,$\\sim$\\,700--900~K to changes in metallicity and surface gravity using a different grid of synthetic spectra by Burrows et al. (\\cite{burrows06}). \\sori~is brighter at $K$ relative to $J$ or $H$ by a factor of $\\sim$1.4. Table~5 and Fig.~4 of Liu et al. (\\cite{liu07}) suggest that log\\,$g$ is thus lower than the field by about 1.0\\,dex, in agreement with previous determinations. We caution that current state-of-the-art synthetic spectra do not provide detailed fits to the observed data (e.g., Burrows et al. \\cite{burrows06}). Therefore, any quantitative result derived from the direct comparison of observations to models awaits further confirmation. On the contrary, for a given temperature qualitative predictions on the atmospheric relative behavior as a function of gravity and metal content can be more reliable. Metallicity might also be an issue. The \\so~cluster has solar abundance ([Fe/H]\\,=\\,0.0\\,$\\pm$\\,0.1~dex; Caballero \\cite{caballero06}); we do not expect any metallicity effect when comparing cluster members to the field. From theoretical considerations, the effects of increasing abundance and decreasing gravity on the near-infrared spectra of cool T dwarfs are similar. From Fig.~20 of Burrows et al. (\\cite{burrows06}), which shows $(J-K)$ against $T_{\\rm eff}$ for various gravities and metal abundances, we infer a metallicity of [Fe/H]\\,$\\sim$\\,$+$0.5 dex for \\sori~if its reddish effect was all due to metallicity. A similar rich metal content is obtained from Liu et al. (\\cite{liu07}). Nevertheless, the super-solar metallicity explanation, although possible, seems unlikely. On the one hand, the metallicity distribution of F, G, and K dwarf stars in the solar neighborhood peaks at around [Fe/H]\\,=\\,0.0 dex and roughly extends up to $+$0.5~dex; less than $\\sim$10\\%~of the stars are more metal-rich than $+$0.3 dex (Valenti et al. \\cite{valenti05}; Santos et al. \\cite{santos05}; Boone et al. \\cite{boone06}). On the other hand, the IRAC/Spitzer data of the S\\,Ori object do not support the high metallicity case. Burgasser et al. (\\cite{burgasser06b}) and Liebert \\& Burgasser (\\cite{liebert07}) have demonstrated 2MASS\\,J12373919$+$6526148 (T6.5) and 2MASS J09373487$+$2931409 (T6p) to be old, high surface gravity brown dwarfs with sub-solar abundance. These field dwarfs display [3.6]\\,$-$\\,[4.5] colors slightly redder (by 0.15\\,mag) than expected for their assigned spectral types. In contrast to what could be inferred from the near-infrared colors, this would indicate that \\sori~is a low metallicity T dwarf. Thus, a low-gravity atmosphere remains as the most likely explanation to account for the observed photometry of \\sori. \\begin{figure*} \\centering \\includegraphics[width=9cm]{4aanda.ps}~~~ \\includegraphics[width=9cm]{5aanda.ps} \\caption{Color-magnitude diagram of \\so~low mass members. \\sori~is labeled. The 3-Myr isochrone by Chabrier \\& Baraffe (\\cite{chabrier00}) is overplotted onto the data as a solid line in the left panel (see text for $T_{\\rm eff}$, log\\,$L/L_\\odot$ conversion into observables). The right panel shows the sequence of M5.5--T8 field dwarfs at the distance of the cluster (solid line, spectral types are indicated). The dotted line stands for the field sequence shifted by $-$1.8 mag in the $J$-band to match the photometric trend delineated by \\so~members. Objects with $J$\\,$-$\\,[3.6] colors significantly redder than the cluster sequence show infrared flux excesses likely due to circum(sub)stellar disks (Caballero et al. \\cite{caballero07}).} \\label{colmag} \\end{figure*} No obvious infrared flux excesses are detectable in the IRAC/Spitzer [3.6] and [4.5] bands, suggesting that there is no envelope or disk around \\sori~emitting intensively at these wavelengths. A possitive detection would have provided strong evidence for its youth. However, we remark that disks around young, low-mass brown dwarfs (close to the deuterium burning mass limit) are seen at wavelengths longer than 5\\,$\\mu$m (Luhman et al. \\cite{luhman05}; Caballero et al. \\cite{caballero07}; Zapatero Osorio et al. \\cite{osorio07}), while the observed fluxes in the near-infrared up to 5\\,$\\mu$m are photospheric in origin. The public [5.8]- and [8.0]-band IRAC/Spitzer images are not conclusive for \\sori. \\subsection{Color-magnitude diagram} The photometric sequence of \\so~substellar members, including \\sori, is shown in the $J$ vs $J$\\,$-$\\,[3.6] color-magnitude diagram of Fig.~\\ref{colmag} (Caballero et al. \\cite{caballero07}; Zapatero Osorio et al. \\cite{osorio07}). This sequence follows a relatively smooth progression with increasing color down to $J$\\,$\\sim$20 and $J$\\,$-$\\,[3.6]\\,=\\,2.8~mag. The location of \\sori~suggests that the $J$\\,$-$\\,[3.6] index suddenly turns toward bluer values at a nearly unchanged $J$ magnitude. A similar turnover (ocurring at spectral types L7--L8, 1400--1300~K) is also observed in field ultracool dwarfs. The field sequence of objects with spectral types M5.5--T8 moved to the distance of the \\so~cluster is displayed in the right panel of Fig.~\\ref{colmag} (absolute magnitudes and colors are adopted from Patten et al. \\cite{patten06}, and references therein). Because of their very young age, \\so~low-mass stars and substellar objects are in the phase of gravitational contraction (e.g., Chabrier \\& Baraffe \\cite{chabrier00}). Cluster members thus show larger size and luminosity than their older counterparts of related colors in the field. As seen in Fig.~\\ref{colmag}, the average cluster photometric sequence appears brighter than the field by about 1.8~mag in the $J$-band. The dotted line in Fig.~\\ref{colmag} (right panel) represents the field sequence normalized to the \\so~locus of late-M and early-L cluster members. \\sori~nicely sits on the location expected for \\so~T-type members. We have also compared our data to the {\\sc cond} and {\\sc dusty} solar metallicity evolutionary models by Chabrier \\& Baraffe (\\cite{chabrier00}). For the age range 1--8\\,Myr, substellar objects with $T_{\\rm eff}$ between 700 and 1300~K, corresponding to the mass interval $\\sim$1--6\\,\\mj, show log\\,$g$\\,=\\,3.0--4.0 dex, which coincides within the large error bar with the surface gravity estimation for \\sori. The 3-Myr isochrone (Chabrier \\& Baraffe \\cite{chabrier00}) is displayed along the photometric sequence of \\so~in the left panel of Fig.~\\ref{colmag}. Theoretical surface temperatures and luminosities were converted into observed magnitudes and colors using the color-temperature-spectral type and spectral type-bolometric correction relationships given in the literature (Dahn et al. \\cite{dahn02}; Vrba et al. \\cite{vrba04}; Knapp et al. \\cite{knapp04}; Patten et al. \\cite{patten06}). The model convincingly reproduces the cluster low mass sequence except for the fact that \\sori~appears overluminous by about 1~mag, which might suggest binarity. This was also discussed in Zapatero Osorio et al. (\\cite{osorio02a}). However, there are issues that prevent us from concluding whether this object is double or whether models make wrong predictions for the smallest masses and young ages: {\\sl (i)} as seen from the field sequence (Fig.~\\ref{colmag}, right panel), there is a $J$-band brightening across the color turnover that theory fails to reproduce (Vrba et al. \\cite{vrba04}; Knapp et al. \\cite{knapp04}). {\\sl (ii)} The blue color turnover takes place at a roughly constant temperature in the field ($\\sim$1300--1400~K), and all color-temperature-bolometric correction transformations show a sharp change at this point; on the contrary, the evolutionary models available to us do not have a complete temperature sampling (this may explain the abrupt color reversal at the bottom of the isochrone in Fig.~\\ref{colmag}). {\\sl (iii)} The relations used to transform theoretical predictions into observables are obtained for high-gravity field objects. It is now known that gravity impacts significantly the near- and mid-infrared colors of T dwarfs (Leggett et al. \\cite{leggett07}; Burrows et al. \\cite{burrows06}), whereas the colors of the warmer M and L types are not so sensitive to the gravity parameter. Transformations are thus expected to be gravity-dependent for the coolest temperatures. It becomes necessary to find a physical explanation for the $J$ brightening feature and to discover more \\so~T-type, planetary-mass members for a proper comparison with evolutionary tracks." }, "0710/0710.2295_arXiv.txt": { "abstract": "I discuss the development and resolution of the solar neutrino problem, as well as opportunities now open to us to extend our knowledge of main-sequence stellar evolution and neutrino astrophysics. ", "introduction": "} This paper is based on a talk given at the Caltech conference \\cite{caltech} ``Nuclear Astrophysics 1957-2007: Beyond the First 50 Years,\" July 23-27, 2007, which focused on the state of nuclear astrophysics fifty years after the seminal paper of Burbidge, Burbidge, Fowler, and Hoyle \\cite{bbfh}. The quest to measure solar neutrinos, and later to resolve the solar neutrino problem, began in the early days of nuclear astrophysics, with the first efforts to understand proton burning in main sequence stars. I would like to review that history, our current understanding of solar neutrinos, and open questions in neutrino physics, and discuss some opportunities for further solar neutrino measurements. Solar neutrino physics brings together stellar modeling, nuclear reactions, and observation. A key early development was the 1959 Holmgren and Johnston \\cite{holmgren} measurement of the S-factor for the pp-chain reaction $^3$He($\\alpha,\\gamma)^7$Be, which proved to be surprisingly large. This implied that the sun could produce some of its energy through the more temperature dependent ppII and ppIII cycles of the pp chain, elevating the neutrino fluxes expected from $^7$Be electron capture and $^8$B $\\beta$ decay. (See Fig. 1.) These neutrinos contribute to ground- and excited-state transitions in $^{37}$Cl($\\nu$,e)$^{37}$Ar, a reaction for detecting neutrinos that had been proposed by Pontecorvo \\cite{pontecorvo} in 1946 and considered by Alvarez \\cite{alvarez} in 1949 (who studied backgrounds that might inhibit, for example, a reactor neutrino experiment). Alvarez's envisioned experiment -- proposed as a test of the distinguishability of the neutrino and antineutrino, prior to the discovery of parity violation -- was later conducted by Davis at Savannah River \\cite{jandr,davis55}. \\begin{figure} \\begin{center} \\includegraphics[width=12cm]{fig1.pdf} \\end{center} \\caption{The three cycles of the pp chain and associated neutrinos.} \\label{fig1} \\end{figure} The Caltech effort in nuclear astrophysics brought three young researchers, Bahcall, Iben, and Sears, together in the summer of 1962. Stimulated in part by the Holmgren and Johnston measurement, they began construction of a solar model to predict the solar core temperature, the most important parameter governing the competition between the ppI, ppII, and ppIII cycles, and to provide the first quantitative estimate of the resulting neutrino fluxes. The Bahcall, Fowler, Iben, and Sears model, published in 1963, predicted a counting rate for Davis's proposed 100,000-gallon chlorine solar neutrino detector of about one event per day. A key development occurred in 1963 when, during a seminar by Bahcall at Copenhagen, Mottelson inquired about the importance of neutrino excitation of excited states in $^{37}$Ar \\cite{jandr}. In 1964 Bahcall and Barnes \\cite{barnes} pointed out that a calibration of the excited state contribution to the $^{37}$Cl cross section could be made by measuring the delayed protons from the analog $\\beta$ decay of $^{37}$Ca, the isospin mirror of $^{37}$Cl. Effectively the lifetime of $^{37}$Ca would be a test of the elevated cross section predicted on the basis of the excited-state contribution to $^8$B neutrino absorption. Subsequent measurements by Hardy and Verrall \\cite{hardy} and Reeder, Poskanzer, and Esterlund \\cite{poskanzer} established the importance of the excited-state contribution. [These experiments were later repeated in a manner that was kinematically complete: see Adelberger et al. \\cite{adelberger} for a discussion.] Bahcall \\cite{bahcall64} and Davis \\cite{davis64} published companion letters in March 1964 arguing the adequacy and feasibility of a 100,000-gallon Cl experiment to measure solar neutrinos. Excavation of the detector cavity in the Homestake Mine began in summer 1965. The tank was installed and filled by the next year. The first results were announced by Davis, Harmer, and Hoffman \\cite{davis68} in 1968, an upper bound of 3 SNU (1 SNU = 10$^{-36}$ capures/Cl atom/sec), below the standard solar model (SSM) prediction of Bahcall, Bahcall, and Shaviv of 7.5 $\\pm$ 3 SNU \\cite{bbs}. This result and associated theoretical work on suggested solutions led to a series of experiments -- Gallex/GNO/SAGE \\cite{gallex,gno,sage}, Kamiokande \\cite{kamiokande}, Super-Kamiokande \\cite{sk} , and the Sudbury Neutrino Observatory \\cite{sno}. Efforts by Borexino \\cite{borexino} and KamLAND \\cite{kamland} to measure the $^7$Be neutrinos are currently underway or in preparation. These experiments are important as tests of the SSM and of the new neutrino physics that proved to be the source of the solar neutrino problem. ", "conclusions": "Neutrino astrophysics and the theories of the origin of the elements, the main theme of this conference, share a common history. Laboratory astrophysics has made solar neutrino physics into a quantitative field, and allowed experimenters to anticipate the kinds of major discoveries that justified experiments like SNO and Super-Kamiokande. The results -- discovery of neutrino mass and flavor mixing characterized by large angles -- are of great importance, providing our first constraints on physics beyond the SM of particle physics. But as summarized here, the list of remaining laboratory neutrino physics questions is long. The answers to the open questions will be important in helping us characterize extreme astrophysical and cosmological neutrino environments. The needed 20-year program of laboratory and astrophysical neutrino studies is not unlike the laboratory/astrophysics interface that Willie Fowler cultivated to help us understand the origin of the elements. Despite the current focus on particle physics properties of neutrinos, solar neutrino spectroscopy remains an important probe of the SSM and stellar evolution. The arguments for measuring the CNO flux, using our sun as a calibrated laboratory, seem particularly strong. Such a program would effectively test our understanding of the hydrogen burning mechanism for massive main-sequence stars. It would also address the primary discrepancy in the SSM, the tension between helioseismology and neutrino flux predictions that follows from new analyses of surface metallicity." }, "0710/0710.2892_arXiv.txt": { "abstract": "The recently recognized class of ``transitional disk\" systems consists of young stars with optically-thick outer disks but inner disks which are mostly devoid of small dust. Here we introduce a further class of ``pre-transitional disks\" with significant near-infrared excesses which indicate the presence of an optically thick inner disk separated from an optically thick outer disk; thus, the spectral energy distributions of pre-transitional disks suggest the incipient development of disk gaps rather than inner holes. In UX Tau A, our analysis of the {\\it Spitzer} IRS spectrum finds that the near-infrared excess is produced by an inner optically thick disk and a gap of $\\sim$56 AU is present. The {\\it Spitzer} IRS spectrum of LkCa 15 is suggestive of a gap of $\\sim$46 AU, confirming previous millimeter imaging. In addition, UX Tau A contains crystalline silicates in its disk at radii $\\gtrsim$ 56 AU which poses a challenge to our understanding of the production of this crystalline material. In contrast, LkCa 15's silicates are amorphous and pristine. UX Tau A and LkCa 15 increase our knowledge of the diversity of dust clearing in low-mass star formation. ", "introduction": "Previous studies have revealed stars with inner disks that are mostly devoid of small dust, and these ``transitional disks\" have been proposed as the bridge between Class II objects, young stars surrounded by full disks accreting material onto the central star, and Class III objects, stars where the protoplanetary disk is mostly dissipated and accretion has stopped (e.g. Strom et al. 1989, Skrutskie et al. 1990; Stassun et al. 2001). New spectra from the {\\it Spitzer Space Telescope} which greatly improve our resolution in the infrared have been used to define the class of ``transitional disks\" as those with spectral energy distributions (SEDs) characterized by a significant deficit of flux in the near-infrared relative to optically thick full disks, and a substantial infrared excess in the mid- and far-infrared. Extensive modeling studies of several transitional disks around T Tauri stars \\citep{dalessio05, uchida04, calvet05, espaillat07} and F-G stars \\citep{brown07} have been presented. In particular, the SEDs of the transitional disks of the T Tauri stars (TTS) CoKu Tau$/$4 \\citep{dalessio05}, TW Hya \\citep{calvet02, uchida04}, GM Aur, DM Tau \\citep{calvet05}, and CS Cha \\citep{espaillat07} have been explained by modeling the transitional disks with truncated optically thick disks with most of the mid-infrared emission originating in the inner edge or ``wall\" of the truncated disk. In all these cases, except in CoKu Tau$/$4, material is accreting onto the star, so gas remains inside the truncated disk, but with a small or negligible amount of small dust, making these regions optically thin. Here we present models of UX Tau A and LkCa 15, low-mass pre-main sequence stars in the young, $\\sim$1 Myr old Taurus star-forming region which have been previously reported as transitional disks \\citep{furlan06, bergin04}. We present evidence for gaps in optically thick disks, as opposed to ``inner holes\", that is, large reductions of small dust from the star out to an outer optically thick wall. ", "conclusions": "Here we introduce the ``pre-transitional disk\" class where we see the incipient development of disk gaps in optically thick protoplanetary disks as evidenced by significant near-infrared excesses when compared to the Taurus median SED and previously studied transitional disks \\citep{dalessio05, calvet02, uchida04, calvet05, espaillat07}. The pre-transitional disk of UX Tau A has a $\\sim$56 AU gap as opposed to an inner hole. It is also possible to fit LkCa 15's SED with a $\\sim$46 AU gap that contains some optically thin dust; a model that has a hole rather than a gap also fits its SED and future near-infrared interferometry may be able to discriminate between these models. However, the truncation of LkCa 15's outer disk at $\\sim$46 AU is consistent with resolved millimeter interferometric observations \\citep{pietu06} which makes it one of three inner disk holes imaged in the millimeter (TW Hya: Hughes et al. 2007; GM Aur: Wilner et al. in preparation). In addition to our sample, the disks around F-G stars studied by \\citet{brown07} also belong to the pre-transitional disk category. The large gaps that are being detected in pre-transitional disks are most likely due to observational bias since larger gaps will create larger mid-infrared deficits in the SED. Smaller gaps will most likely have less apparent dips in their SEDS and be more difficult to identify, however, if their gaps contain some optically thin material the silicate emission in these objects should be much stronger than can be explained by a full disk model. The existence of an inner optically thick disk may be an indicator of the first stages of disk clearing that will eventually lead to the the inner holes that have been seen in previously reported transitional disks; this has important implications on disk evolution theories since only planet-formation can account for this structure. Hydrodynamical simulations have shown that a newly formed planet could accrete and sweep out the material around it through tidal disturbances and this is sufficient in producing the hole size in CoKu Tau$/$4 \\citep{quillen04}, even maintaining substantial accretion rates \\citep{varniere06}. Moreover, \\citet{najita07} have found that the intrinsic properties of transitional disks may favor planet formation. Another proposed formation mechanism for the holes in transitional disks is photoevaporation, in which a photoevaporative wind halts mass accretion towards the inner disk and material in this inner disk is rapidly evacuated creating an inner hole \\citep{clarke01}; the hole then increases in size as the edge continues photoevaporating \\citep{alexanderarmitage}. Neither this model nor the inside-out evacuation induced by the MRI \\citep{chiang07} would explain how an optically thick disk accreting at a sizable accretion rate (see Table 1) would remain inside the hole. Rapid dust growth and settling has also been proposed to explain the holes in disks \\citep{lin04}. Again, this does not account for the presence of optically thick inner disk material given that theory suggests grain growth should be fastest in the inner disk, not at some intermediate radius \\citep{weiden97, chiang99}. Our sample also has interesting dust compositions (Watson et al. 2007, Sargent et al. in preparation). LkCa 15 has an amorphous silicate feature indicating little if any processing leading to the crystallization seen in other young stars. Amorphous silicates are also seen in CoKu Tau$/4$, DM Tau, and GM Aur. In contrast, UX Tau A is different from all the other transitional disks because it has crystalline silicate emission features in addition to amorphous silicate emission features (Fig. 1 inset). The wall at $\\sim$56 AU is the main contributor to the crystalline silicate emission since it dominates the flux in the mid- and far-infrared. This raises the question of whether crystalline silicates are created close to the star or if they can be created in situ at $\\sim$56 AU. If the former, it challenges current radial-mixing theories, none of which can get significant amounts of crystalline silicates out to this distance \\citep{gail01, kellergail04}. One possibility for {\\it in situ} processing may be collisions of larger bodies, which might produce small grains heated sufficiently to create crystals (S. Kenyon, personal communication). Pre-transitional disks offer further insight into the diversity of the ``transitional disk\" class and future studies of these disks will greatly advance our understanding of disk evolution and planet formation. \\vskip -0.1in" }, "0710/0710.0545_arXiv.txt": { "abstract": "We present a model to estimate the synchrotron radio emission generated in microquasar (MQ) jets due to secondary pairs created via decay of charged pions produced in proton-proton collisions between stellar wind ions and jet relativistic protons. Signatures of electrons/positrons are obtained from consistent particle energy distributions that take into account energy losses due to synchrotron and inverse Compton (IC) processes, as well as adiabatic expansion. The space parameter for the model is explored and the corresponding spectral energy distributions (SEDs) are presented. We conclude that secondary leptonic emission represents a significant though hardly dominant contribution to the total radio emission in MQs, with observational consequences that can be used to test some still unknown processes occurring in these objects as well as the nature of the matter outflowing in their jets. ", "introduction": "\\label{intro} X-ray binary systems (XRBs) are composed by either a stellar mass black hole or a neutron star, and a normal (non degenerated) star which supplies matter to the compact object through the formation of an accretion disk. Some ~260 XRBs are known up to now \\cite{liu06} probably corresponding to an underlying population of some tens of thousands of compact objects in our Galaxy. A few of these sources present also non-thermal radio emission, hence evidencing the existence of mechanism(s) capable of injecting and/or accelerating large numbers of relativistic particles. Some radio emitting X-ray binary systems (REXBs) have been observed showing ejection of material at relativistic velocities and to display jets like those seen in quasars and active galactic nuclei but at $\\sim 10^{-6}$ times shorter scales. This analogy is the reason for calling them microquasars (MQs) \\cite{mirabel99} and making them some of the most interesting objects for astrophysics. Furthermore, attention on these objects has grown since the proposal of Paredes et al. (2000) \\cite{paredes00} of MQs as counterparts of some of the unidentified gamma-ray sources of the EGRET catalog \\cite{hartman99} and hence pointing them as plausible high energy emitters. A strong confirmation of this association has come from the detections of the MQs LS 5039 and LS I +61 303 at Tev energies using respectively the ground-based Cherenkov telescopes HESS \\cite{aharonian05} and MAGIC \\cite{albert06}, giving support and empowering at the same time a number of previous detailed studies centered on the mechanisms operating in these sources in order to explain the gamma ray domain (see, e.g., \\cite{bosch05} and \\cite{romero05}). Moreover, a jet origin of the emission from MQs has been suggested from the observation of syncrothron emission of relativistic electrons/positrons extending from the radio all the way into the X-ray regime. In this sense jet-like models have focused on different approaches regarding the particle origin that could generate the required emission properties in a consistent way. Some of them consider leptons directly injected at the base of the jet and, in extending outwards, Compton-interaction with external/self-created photon fields produces high energy radiation. Other models deal with an hadronic origin of the high energy emission, through proton-proton interactions and pion decay producing gamma rays and leaving the resulting co-generated leptons as low energy emitters. The present work refers to the later procedure, focusing on the modelisation of the secondary leptonic synchrotron emission in order to constrain the characterization of MQ jets. An outline of the model is given in the next, followed by the results showing the SEDs and lightcurves under different parameter assumptions and the conclusions that can be extracted from them. ", "conclusions": "SEDs are obtained for different magnetic field values, electron/positron spectral indices and spatially distributed disks. We have estimated also the expected emission along the jet at 1 and 5 GHz. Leptons are injected in the context of hadronic secondaries generation within a detailed model that takes into account in a consistent way particle injection mechanisms and cooling due to radiation processes and adiabatic expansion. The luminosities obtained are slightly lower than in the models based on primary leptons injection, and must be considered complementary to them. However, we note that within our model there is no requirements of acceleration processes along the jet to obtain the final emission results. Such acceleration processes are still not well understood, although. They could come from diffusive shock acceleration along the jet when fresh ejecta interact with previous blobs of plasma already outflowing at lower velocities. Other scenarios assume a continuous energy transfer mechanism from the magnetic field to the matter content of the jet in such a way that the resulting parsec-scale radio emission can be explained. The fact of studying alternative models were particles are directly injected until a certain height along the jet can constrain the amount of acceleration required and contribute to the understanding of the physical mechanisms that can lead to such processes. Signatures at different distances along the jet and specific spectral features detectable for reasonable parameter values treated in our numerical simulations have the potential to be an important clue for determining the matter content of jets. In particluar, highly resolved observations at 1 and 5 GHz could determine if leptons are present at heights $10^{12-13}$ cm at the edge of the binary system typical region where wind matter from the companion is still significant. If electrons/positrons still show high energies due to a recent injection from hadronic interactions at these parts of the jet, it could be a signature of secondary generation without the necessity of invoking additional acceleration processes." }, "0710/0710.5010_arXiv.txt": { "abstract": "Recent observations of XTE J1739-285 suggest that it contains a neutron star rotating at 1122 Hz\\cite{Kaaret2007}. Such rotational frequency would be the first for which the effects of rotation are significant. We study the consequences of very fast rotating neutron stars for the potentially observable quantities as stellar mass and pulsar period. ", "introduction": "\\label{sect:introd} Neutron stars with their very strong gravity can be very fast rotators. Theoretical studies show that they could rotate at sub-millisecond periods, i.e., at frequency $f=1/{\\rm period}>$1000 Hz\\cite{CST1994,Salgado1994}. The first millisecond pulsar B1937+21, rotating at $f=641$ Hz\\cite{Backer1982}, remained the most rapid one during 24 years after its detection. In January 2006, discovery of a more rapid pulsar J1748-2446ad rotating at $f=716$ Hz was announced \\citep{Hessels2006}. However, such sub-kHz frequencies are still too low to significantly affect the structure of neutron stars with $M>1M_\\odot$ \\cite{STW1983}. Actually, they belong to a {\\it slow rotation} regime, because their $f$ is significantly smaller than the mass shedding (Keplerian) frequency $f_{\\rm K}$. Effects of rotation on neutron star structure are then $\\propto (f/f_{\\rm K})^2\\ll 1$. Rapid rotation regime for $M>1M_\\odot$ requires submillisecond pulsars with supra-kHz frequencies $f>1000$ Hz. Very recently Kaaret et al.\\cite{Kaaret2007} reported a discovery of oscillation frequency $f=1122$ Hz in an X-ray burst from the X-ray transient, XTE J1739-285. According to Kaaret et al.\\cite{Kaaret2007} \"this oscillation frequency suggests that XTE J1739-285 contains the fastest rotating neutron star yet found\". If confirmed, this would be the first detection of a sub-millisecond pulsar (discovery of a 0.5 ms pulsar in SN1987A remnant announced in January 1989 was withdrawn one year later). Rotation at $f>1000$ Hz is sensitive to the stellar mass and to the equation of state (EOS). Hydrostatic, stationary configurations of neutron stars rotating at given rotation frequency $f$ form a one-parameter family, labeled by the central density. This family - a curve in the mass - equatorial radius plane - is limited by two instabilities. On the high central density side, it is instability with respect to axi-symmetric perturbations, making the star collapse into a Kerr black hole. The low central density boundary results from the mass shedding from the equator. In the present paper we show how rotation at $f>1000$ Hz is sensitive to the EOS, and what constraints on the EOS of neutron stars result from future observations of stably rotating sub-millisecond pulsars. ", "conclusions": "The $M(R_{\\rm eq})$ curve for $f\\gtrsim 1400$ Hz is flat. Therefore, for given EOS the mass of NS is quite well defined. Conversely, measured mass of a NS rotating at $f\\gtrsim 1400$ Hz will allow us to unveil the actual EOS. The \"Newtonian\" formula for the Keplerian frequency works surprisingly well for precise 2-D simulations and sets a firm upper limit on $R_{\\rm eq}$ for a given $f$. Finally, observation of $f\\gtrsim 1200$ Hz sets stringent limits on the initial mass of the nonrotating star which was spun up to this frequency by accretion. \\label{sect:discuss}" }, "0710/0710.0621_arXiv.txt": { "abstract": "X-ray spectra from stellar coronae are reprocessed by the underlying photosphere through scattering and photoionization events. While reprocessed X-ray spectra reaching a distant observer are at a flux level of only a few percent of that of the corona itself, characteristic lines formed by inner shell photoionization of some abundant elements can be significantly stronger. The emergent photospheric spectra are sensitive to the distance and location of the fluorescing radiation and can provide diagnostics of coronal geometry and abundance. Here we present Monte Carlo simulations of the photospheric K$\\alpha_1,\\alpha_2$ doublet arising from quasi-neutral Fe irradiated by a coronal X-ray source. Fluorescent line strengths have been computed as a function of the height of the radiation source, the temperature of the ionising X-ray spectrum, and the viewing angle. We also illustrate how the fluorescence efficiencies scale with the photospheric metallicity and the Fe abundance. Based on the results we make three comments: (1) fluorescent Fe lines seen from pre-main sequence stars mostly suggest flared disk geometries and/or super-solar disk Fe abundances; (2) the extreme $\\approx 1400$~m\\AA\\ line observed from a flare on V~1486~Ori can be explained entirely by X-ray fluorescence if the flare itself were partially eclipsed by the limb of the star; and (3) the fluorescent Fe line detected by {\\it Swift} during a large flare on II~Peg is consistent with X-ray excitation and does not require a collisional ionisation contribution. There is no convincing evidence supporting the energetically challenging explanation of electron impact excitation for observed stellar Fe~K$\\alpha$ lines. ", "introduction": "\\label{s:intro} It is well-established from surveys of the sky at EUV and X-ray wavelengths that all stars with spectral types later than mid-F, except for giants later than mid-K, possess hot outer atmospheres akin to that of the Sun \\citep[e.g.][]{Vaiana.etal:81, Schmitt:97}. While much observational and theoretical effort has been devoted to understanding solar coronal spectra and, in more recent years, toward understanding stellar coronal emission and spectra, comparatively little attention has been devoted to the reprocessing and line fluorescence resulting from this coronal emission by the underlying solar and stellar photospheres. In contrast, considerable effort has been spent on the study of X-ray reprocessing by ``cold'' gas in much more complex systems with more prominent fluorescent features but more uncertain geometries and physical conditions, such as the accretion disks around black holes and non-degenerate objects in X-ray binaries \\citep[e.g.][]{Felsteiner.Opher:76,Hatchett.Weaver:77,Fabian.etal:89, George.Fabian:91,Laor:91, Matt.etal:97, Ballantyne.etal:02, Beckwith.Done:04, Cadez.Calvani:05, Dovciak.etal:04, Laming.Titarchuk:04, Brenneman.Reynolds:06}. The processes involved in photospheric fluorescence by coronal irradiation are the same as those discussed in these works; the main difference here is in the specific geometry of the X-ray source above a quasi-neutral sphere and of the extended, shell-like nature of the coronal source above the photosphere during quiescent conditions. X-rays emitted from a hot ($T\\ga 10^6$~K) corona incident on the underlying photosphere can undergo either Compton scattering or photoabsorption events through the ionization of atoms or weakly ionized species. Through scattering events, photons can be reflected back in a direction towards the stellar surface where they have a finite chance of escape. Compton scattering redistributes the spectrum to lower energies by $\\sim E^2/m_ec^2$ per collision, where $E$ is the photon energy and $m_e$ the electron rest mass. The spectrum reflected from a stellar surface by scattering events is then shifted and broadened towards lower energies. Photoionization events involving X-ray photons directed toward the photosphere are predominantly inner shell interactions with astrophysically abundant elements, the outer and valance cross-sections being very small at these energies. Observable fluorescent lines can then arise as a result of the finite escape probabilities of photons emitted in outward directions by hole transitions in these atoms photoionized in their inner shells. These processes have been described in the solar context by, e.g., \\citet{Tomblin:72} and \\citet{Bai:79} (B79), and more recently for arbitrarily photoionized slabs by \\citet{Kallman.etal:04}. The strongest of the fluorescent lines for a plasma of approximately cosmic composition is the $2s$-$1p$ 6.4~keV Fe K$\\alpha$ doublet occurring following ejection of a $1s$ electron. It has been observed in solar spectra on numerous occasions \\citep[e.g.][]{Neupert.etal:67, Doschek.etal:71, Feldman.etal:80, Tanaka.etal:84, Parmar.etal:84, Zarro.etal:92}. The mechanism of fluorescence by the thermal X-ray coronal continuum was suggested by \\citet{Neupert.etal:67}, and was firmly established on more theoretical grounds by \\citet{Basko:78,Basko:79} and B79. \\citet{Parmar.etal:84} provided convincing observational confirmation based on flare spectra obtained by the {\\it Solar Maximum Mission}, though it has also been noted that contributions from non-thermal electron impact might also be present during hard X-ray bursts \\citep*[e.g.][]{Emslie.etal:86, Zarro.etal:92}. B79 pointed out that, for a given source spectrum, the observed flux of Fe~K$\\alpha$ photons from the photosphere depends on essentially three parameters: the photospheric iron abundance; the height of the emitting source; and the heliocentric angle between the emitting source and observer. \\citet{Phillips.etal:94} used Fe K$\\beta$ observations to probe the difference between the photospheric and coronal iron abundance for flares observed by the {\\it Yohkoh} satellite. More importantly for the stellar case, the spatial aspects of photospheric fluorescent line formation suggest its application to understanding the spatial distribution of coronal structures and flares on stars of different spectral type and activity level to the Sun \\citep{Drake.etal:99}. Indeed, Fe~K fluorescence has recently been detected during flares on the active binary II~Peg \\citep{Osten.etal:07} and on the single giant HR~9024 \\citep{Testa.etal:07}. The Fe~K line has also been seen in a growing sample of pre-main sequence (PMS) stars \\citep{Imanishi.etal:01,Favata.etal:05,Tsujimoto.etal:05, Giardino.etal:07}, in which the line is thought to originate predominantly from the irradiated protoplanetary disk rather than the photosphere. Since the work of \\cite{Bai:79}, there have been no concerted efforts to extend models of photospheric fluorescence for coronal excitation sources with other characteristics. Fluorescent lines other than Fe~K$\\alpha$ have also not yet, to our knowledge, been studied by other workers in any detail in this context. A reasonably strong feature observed in solar spectra near 17.62~\\AA\\ had been identified with Fe L$\\alpha$ photospheric fluorescence \\citep{McKenzie.etal:80, Phillips.etal:82}, but calculations of the expected line strength was shown by \\citet{Drake.etal:99} to be much too weak to explain the feature, and these authors instead identified the line with a transition in Fe~XVIII arising from configurational mixing and both seen in {\\it Electron Beam Ion Trap} spectra and predicted by theory \\citep{Cornille.etal:92}. However, given the potential diagnostic value of photospheric fluorescence, other lines, such as O~K$\\alpha$, are possibly observable with very high quality observations and warrant further study. The capabilities of current X-ray missions such as {\\it Chandra}, {\\it XMM-Newton}, {\\it Swift} and {\\it Suzaku} to detect fluorescent lines further motivates a re-examination of the photospheric fluorescence problem in the context of stellar coronae, photospheres and protoplanetary disks. We restrict the study in hand to Fe~K fluorescence from stellar photospheres and defer detailed discussions of protoplanetary disks and fluorescence from other elements to future work. ", "conclusions": "The main scientific motivation for this work is to provide the foundation to use Fe fluorescence as a quantitative diagnostic of coronal and flare geometry. There now exists a handful of detections of fluorescent emission from stars. Sensitivity is currently limited to a large extent by the low spectral resolution of available instruments and progress is expected to accelerate dramatically with the future availability of X-ray calorimeters. Since modern X-ray spectral analyses based on low-resolution CCD pulse-height spectra tend to express line strengths in terms of the line equivalent width, we have computed this quantity for the case of $\\theta=0$ and the ranges of heights and X-ray temperatures investigated in \\S\\ref{s:newcalcs}. The equivalent width in this context refers to the fluorescent, processed line seen on top of the continuum of the ionising coronal spectrum. While the $\\theta=0$ case gives the most optimistic line strength, we note that $f(\\theta)$ is quite slowly varying for angles $\\theta \\la 45^\\circ$ for the flare heights for which significant Fe\\,K${\\alpha}$ might be observed. The equivalent widths are illustrated in Figure~\\ref{f:ew} and listed in Table~\\ref{t:fekaEW1}. \\subsection{Fluorescence from Pre-Main Sequence Stars} The observability of the cold Fe~K$\\alpha$ line is of course strongly dependent on the quality of the X-ray spectrum obtained. The most extensive study of PMS Fe fluorescence to date is that based on Chandra observations of the Orion Nebula Cluster by \\citet{Tsujimoto.etal:05}. This study detected significant 6.4~keV excesses attributable to Fe fluorescence for 7 out of 127 sources found to have significant counts in the 6-9~keV band. Equivalent widths were in the range 110-270~eV at plasma temperatures of $\\sim 3$-10~keV. There is clearly a strong selection effect here and these fluorescent line strengths likely represent the upper end of the distribution. Our calculations for a flare at scale height $h=0 R_\\star$ are also appropriate for a flare occurring above an infinite plane, such as might approximate a disk-encircled PMS star. As in the photospheric case, the fluorescence problem can be treated orthogonally from the ionisation structure of the disk, which is not greatly altered from its very largely neutral overall state by X-rays from a typical flare. Any small degree of X-ray photoionisation will also not affect Fe~K$\\alpha$ line strengths because fluorescence yields are essentially invariant for lower Fe ions. Our calculations indicate that attaining equivalent widths much in excess of 100~eV is not straightforward for such a simple geometry for the plasma temperatures observed in the fluorescing X-ray spectra. This finding is in agreement with earlier calculations by \\citet{Matt.etal:91} and \\citet{George.Fabian:91}, who find equivalent widths of $\\sim 150$~eV for a flat disk illuminated by X-rays with power-law photon spectral energy distributions. There are at least four ways in which equivalent widths might be elevated above the values we find: (1) super-solar Fe abundance in the disk material, possibly arising as a result of an elevated dust-to-gas ratio; (2) disk flaring, resulting in a solid angle coverage $> 2\\pi$; (3) line-of-sight obscuration of the central flaring source (but not fluorescent line photons) by optically thick structures such as the star itself (ie the flare occurring on the far hemisphere); and (4) fluorescence contributions from ionisation by non-thermal electrons. By analogy with the solar case, in which excitation by non-thermal electrons is usually negligible \\citep{Parmar.etal:84,Emslie.etal:86} and is much more difficult on energetic grounds, we consider (4) the {\\em least} plausible of these. \\citet{Ballantyne.Fabian:03} have also shown in the accretion disk context that Fe~K production by non-thermal electron bombardment requires 2-4 orders of magnitude greater energy dissipation in the electron beam than is required for an X-ray photoionization source. Disk flaring can give rise to increased line strengths by factors $< 2$ simply from consideration of the increased solid angle coverage possible compared with an infinite flat disk. It is also difficult to envisage enhanced Fe abundances in the disk being able to elevate line strengths by more than a factor of a few. Of some interest, then, is the observation of an Fe~K$\\alpha$ line equivalent width of $\\approx1400$~m\\AA\\ during the rise phase of a flare on the PMS Orion nebula star forming region object V~1486~Ori by \\citet{Czesla.Schmitt:07}. Such an enhancement over an infinite disk value of $\\sim 150$~m\\AA , even with a large degree of disk flaring, would still require extreme enhancements of the disk Fe abundance by an order of magnitude or more \\citep[e.g. \\S\\ref{s:fesens} and][]{Matt.etal:97} were the line due to photoionisation by the {\\em directly observed} continuum. We point out, however, that fluorescent line photons from a PMS disk can still be observed when the X-ray flaring source is located behind the star and obscured from the line-of-sight. In the case of the V~1486~Ori flare, the large observed Fe~K$\\alpha$ equivalent width is simply and plausibly associated with a partially obscured flare whose rise phase was not fully observed directly owing to line-of-sight obscuration by the star itself. Such obscured flares will inevitably be the cause of some fraction of observed Fe~K$\\alpha$ lines from PMS stellar disks. This explanation is also more consistent than one relying on preferential disk geometry and Fe abundance with the non-detection of Fe~K$\\alpha$ from a second less extreme flare whose impulsive phase instead appears to have been quite visible. \\subsection{Fluorescence from stellar photospheres} The strong flare in II~Peg observed by {\\it Swift} and analysed by \\citet{Osten.etal:07} presents another interesting case. Photospheric Fe~K$\\alpha$ was clearly detected throughout the event. Equivalent widths for different times in the flare ranged from 18 to 61~eV, with uncertainties in the 20-45\\%\\ range. The authors favoured a collisional excitation mechanism for the line, arguing that fluorescence would be unlikely to produce an observable feature. This assessment employed a simple analytical formula applicable to optically-thin cases in which only a minor fraction of the incident X-ray flux is subject to photoabsorption or scattering \\citep[see e.g.][]{Liedahl:99,Krolik.Kallman:87}. \\citet{Osten.etal:07} correctly noted that the path length required to obtain the observed equivalent widths under such an approximation was similar to the $\\tau=1$ Compton scattering depth, but discounted fluorescence as a possibility on these grounds. Other than the inapplicability of the optically-thin formula for the photospheric fluorescence case, one reason such an argument is overly pessimistic is that incidence angles on the photosphere range from $\\sim 0$--$90^\\circ$ for small scale heights and path lengths for escape are less than penetration depths by the factor of the inverse cosine of these angles. We defer a more detailed treatment of the event to future work, but note here that equivalent widths of 50~eV are achieved for flare heights up to $\\sim 0.2 R_\\star$ or so for the $10^8$~K model in our grid, a temperature similar to the average of the values found for the flare by \\citet{Osten.etal:07}. While collisional ionisation cannot be ruled out observationally as the source of the observed Fe fluorescence, it is not a requirement." }, "0710/0710.5374_arXiv.txt": { "abstract": "While isolated neutron stars (INSs) are among the brightest $\\gamma$-ray sources, they are among the faintest ones in the optical, and their study is a challenging task which require the most powerful telescopes. \\hst\\ has lead neutron star optical astronomy yielding nearly all the identifications achieved since the early 1990s. Here, the major \\hst\\ contributions in the optical studies of INSs and their relevance for neutron stars' astronomy are reviewed. ", "introduction": "\\label{sec:1} Before the launch of \\hst, optical studies of INSs were the exception. In the first 20 years since the pulsars discovery, only the Crab and Vela pulsars were identified (Cocke 1969; Lasker 1976), while optical pulsations were detected from an unidentified source at the center of SNR B0540-69 in the LMC (Middleditch \\& Pennypacker 1985), and only a candidate counterpart was found for the misterious $\\gamma$-ray source Geminga (Bignami et al. 1987). This score was expected to be considerably improved by \\hst, thanks to its much larger sensitivity with respect to ground based telescopes, and to the sharp spatial resolution of the \\wfpc\\ as well as to the near-UV view of the ESA's \\foc. Unfortunately, the spherical aberration of the \\hst\\ optics affected the execution of most approved proposals, except for those aimed at the brightest targets. So, in the early 1990s the leadership in the INSs optical astronomy was still in the hands of ground-based observatories, mainly in those of the ESO \\ntt\\ which secured the identification of Geminga through the proper motion of its counterpart, a technique soon become the standard one, and the likely identifications of the optical pulsar in SNR B0540-69 and of PSR B0656+14 (see Mignani et al. 2000). However, the refurbishment of \\hst\\ in SM-1 (Dec. 2003) brought its performance back to the original expectations and gave it a leading role in INSs' optical astronomy, mantained even after the advent of the 10-m class telescopes. Since \\hst\\ has provided 8 new INSs identifications, against the 2 of the \\vlt\\ and the \\keck\\ (see Mignani et al. 2004), boosting the identification rate by a factor 4. This could have been higher if not for the \\foc\\ removal in SM-3B (March 2002) and for the \\stis\\ failure (Aug. 2004), which alone have yielded nearly all the \\hst\\ INSs identifications, depriving the telescope of its near-UV view. Thus, \\hst\\ observations have opened wide a new, important observing window on INSs and triggered the interest of a larger and larger fraction of the neutron star community. ", "conclusions": "" }, "0710/0710.2248_arXiv.txt": { "abstract": "{} {We intend to establish the X-ray properties of Swift J0732.5--1331 and therefore confirm its status as an intermediate polar.} {We analysed 36\\,240~s of X-ray data from {\\em RXTE}. Frequency analysis was used to constrain temporal variations and spectral analysis used to characterise the emission and absorption properties.} {The X-ray spin period is confirmed to be 512.4(3)~s with a strong first harmonic. No modulation is detected at the candidate orbital period of 5.6~h, but a coherent modulation is present at the candidate 11.3~h period. The spectrum is consistent with a 37~keV bremsstrahlung continuum with an iron line at 6.4~keV absorbed by an equivalent hydrogen column density of around $10^{22}$ atoms~cm$^{-2}$.} {Swift J0732--1331 is confirmed to be an intermediate polar.} ", "introduction": "Intermediate polars (IPs) are a sub-class of cataclysmic variables (CVs). They fill the phase space, in terms of magnetic field strength, and spin and orbital periods, between non-magnetic CVs and the strongly magnetic synchronously rotating polars. The magnetic field strength is believed to be in the range of a few MG to tens of MG at the white dwarf surface. This is large enough to dramatically alter the accretion flow, yet not large enough to synchronize the spin and orbital periods. This magnetic field gives rise to the defining characteristic of the sub-class, that of X-ray variation pulsed at the spin period of the white dwarf. For an exhaustive review of CVs see e.g. \\cite{warner95}. There are between twenty six\\footnote{http://asd.gsfc.nasa.gov/Koji.Mukai/iphome/iphome.html as of 23/8/7.} and fifty IPs currently known (depending on the selection criteria used). The hard X-ray selected object, Swift~J0732.5--1331 (hereafter J0732), is a suspected IP in need of confirmation. The circumstance of its discovery makes J0732 similar to the host of candidate IPs that have been discovered to be powerful emitters of hard X-rays/soft gamma-rays in the 20--100~keV range in the {\\sl INTEGRAL\\/}/IBIS survey \\citep{barlow06}. We have embarked on a campaign of pointed {\\sl RXTE\\/} observations of these hard X-ray discovered candidate IPs. Here we present the first results of our campaign on J0732. ", "conclusions": "The unambiguous X-ray spin period detection at 512.4(3)~s, along with the spectral fit to an absorbed 37~keV bremsstrahlung model with an iron line, confirm the intermediate polar status of Swift J0732.5--1331. We are unable to determine the orbital period from these {\\em RXTE} data although there is some indication of modulation at the previously suggested photometric period of 11.3~h and none at the spectroscopic period of 5.6~h. To conclude we note that this system is similar, in terms of its small $P_{\\rm spin}/P_{\\rm orb}$ value, to the IPs RX~J2133.7+5107 and NY~Lup (IGR~J15479--4529). Both of these are {\\em INTEGRAL} hard X-ray sources and both also have soft X-ray components. We might therefore expect that Swift~J0732.5--1331 would also display such characteristics upon further study." }, "0710/0710.0767_arXiv.txt": { "abstract": "We investigate the possibility that near future observations of ultra-high-energy cosmic rays (UHECRs) can unveil their local source distribution, which reflects the observed local structures if their origins are astrophysical objects. In order to discuss this possibility, we calculate the arrival distribution of UHE protons taking into account their propagation process in intergalactic space i.e. energy losses and deflections by extragalactic magnetic field (EGMF). For a realistic simulation, we construct and adopt a model of a structured EGMF and UHECR source distribution, which reproduce the local structures actually observed around the Milky Way. The arrival distribution is compared statistically to their source distribution using correlation coefficient. We specially find that UHECRs above $10^{19.8}$eV are best indicators to decipher their source distribution within 100 Mpc, and detection of about 500 events on all the sky allows us to unveil the local structure of UHE universe for plausible EGMF strength and the source number density. This number of events can be detected by five years observation by Pierre Auger Observatory. ", "introduction": "\\label{intro} The origin of ultra-high-energy cosmic rays (UHECRs) above $10^{19}$eV is one of the most intriguing problems in astroparticle physics. Akeno Giant Air Shower Array (AGASA) found statistically significant small-scale clusterings of observed UHECR events with large-scale isotropy \\citep*{takeda99}. The AGASA data set of 57 events above $4 \\times 10^{19}$eV contains four doublets and one triplet within separation angle of $2^{\\circ}.5$, consistent with the experimental angular resolution. The chance probability of observing such multiplets under an isotropic distribution is only about 1\\% \\citep*{hayashida00}. A combination of the results of many UHECR experiments (including AGASA) also revealed eight doublets and two triplets within $4^{\\circ}$ on a totally 92 events above $4 \\times 10^{19}$eV \\citep*{uchihori00}. These multiplets suggest that the origins of UHECRs are point-like sources. For identification of UHECR sources, arrival directions of UHECRs have been observed in detail by High Resolution Fly's Eye (HiRes) and Pierre Auger Observatory (Auger). However, so far, these experiments have reported no significant clustering on the arrival distribution above $4 \\times 10^{19}$eV \\citep*{abbasi05,mollerach07}. Recently, several classes of astrophysical objects in many literature has tested for positional correlations with observed arrival directions of UHECRs. The correlations with BL Lac objects were discussed on the assumptions of smaller deflection angles of UHECRs than the experimental angular resolution and/or neutral primaries \\citep*{tinyakov01}, and in consideration with the deflection by Galactic magnetic field (GMF) \\citep*{tinyakov02}. \\cite{gorbunov05} considered various classes of powerful extragalactic sources for the correlation study including small corrections of UHECR arrival directions by GMF. \\cite{hague07} discussed the correlation with nearby active galactic nuclei (AGNs) from RXTE catalog of AGNs. However, these studies have not taken into account UHECR propagation in extragalactic space. UHECRs above $8 \\times 10^{19}$eV lose a significant fraction of their energies by photopion production in collision with the cosmic microwave background (CMB) photons during their propagation \\citep*{berezinsky88,yoshida93}. Thus, UHECRs have {\\it horizons}, which are the maximum distances of their sources that UHECRs can reach the Earth, even if their energies are below $8 \\times 10^{19}$eV at the Earth. The positional correlations between arrival directions of UHECRs and their source candidates outside the horizons are not significant. (In \\cite{hague07}, only nearby AGNs within the horizons are considered.) In addition to the UHECR horizons, deflections due to extragalactic magnetic field (EGMF) are also important since extragalactic cosmic rays are propagated for a much greater distance than in Galactic space. Propagation process of UHECRs should be considered in such correlation studies. \\cite{yoshiguchi03} investigated the correlation between the arrival distribution of UHECRs and their source distribution taking into account UHECR propagation in intergalactic space with a uniform turbulent EGMF whose strength is 1 nG and coherent length is 1 Mpc. The authors adopted a source distribution with $10^{-6}~{\\rm Mpc}^{-3}$ that reproduced the local structures and the AGASA results. They concluded that detection of a few thousand events above $4 \\times 10^{19}$eV reveal observable correlation with the sources within 100 Mpc. However, a uniform turbulent field is not realistic EGMF model. Faraday rotation measurements indicate magnetic field strengths at the $\\mu$G level within inner region ($\\sim$ central Mpc) of galaxy clusters \\citep*{kronberg94}. The evidence for synchrotron emission in numerous galaxy clusters \\citep*{giovannini00} and in a few cases of filaments \\citep*{kim89,bagchi02} also seems to suggest the presence of magnetic fields with $0.1-1.0\\mu$G at cosmological structures. Several numerical simulations of large-scale structure formation have confirmed these magnetic structures \\citep*{sigl03,dolag05}. Based on these studies, in recent years, we calculated propagation of UHE protons in a structured EGMF which well reproduces the local structures actually observed and simulated their arrival distributions with several normalizations of EGMF strength and several number density of UHECR sources \\citep*{takami07}. We constrained the source number density to best reproduce the AGASA results. As a result, $10^{-5}~{\\rm Mpc}^{-3}$ is the most appropriate number density, which is weakly dependent on EGMF strength. (In rectilinear propagation, similar number density is also obtained in \\cite{blasi04,kachelriess05}) However, this has large uncertainty due to the small number of observed events at present. $10^{-4}~{\\rm Mpc}^{-3}$ and $10^{-6}~{\\rm Mpc}^{-3}$ are also statistically allowed. Therefore, it is useful to deliberate the correlation between the arrival distribution and the source distribution in the case of these number densities. Note that we revealed in the paper that this uncertainty will be solved by future increase of detected events. In this study, we calculate the arrival distribution of UHECRs, taking their propagation process into account, and investigate the correlation the arrival distribution and their source distribution in the future. A structured EGMF model and source distribution which can reproduce the local universe actually observed are adopted. The source number density and the EGMF strength are treated as parameters since these have some uncertainty. A goal of this study is that we understand the number of observed events to start to observe the UHECR source distribution by UHECRs and how much the correlation is expected in the future. Auger has already detected more events above $10^{19}$eV than those observed by AGASA \\citep*{roth07}. Nevertheless, the event clustering has not observed, as mentioned above. It might be due to EGMF and/or GMF strong enough not to generate the multiplets or statistical fluctuation for small number of observed events at highest energies. In any case, we should predict and discuss how the arrival distribution reflects UHECR source distribution. Chemical composition of UHECRs is very important for the correlation. If UHECRs are heavier components, magnetic deflections are larger and the correlation is worse. One of observables for study of UHECR composition is the depth of shower maximum, $X_{\\rm max}$, which can be measured by fluorescence detectors. Its average value $\\left< X_{\\rm max} \\right>$ is dependent on UHECR composition and energy. HiRes reported that composition of cosmic rays above $10^{19}$eV is dominated by protons as a result of $X_{\\rm max}$ measurement \\citep*{abbasi05b}. Recent result by Auger is compatible to the HiRes result within systematic uncertainties \\citep{unger07}. However, they concluded that the interpretation of $\\left< X_{\\rm max} \\right>$ distribution is ambiguous because of the uncertainties of hadronic interaction at highest energies. Thus, UHECR composition at highest energies is controversial at present. Despite that, in this study, we assume that all UHECRs are protons since composition above $10^{19}$eV has proton-like feature. This paper is organized as follows: in section \\ref{model} we provide our models of UHECR source distribution and a structured EGMF. In section \\ref{method} we explain our calculation method for the arrival distribution with UHECR propagation and statistical method. In section \\ref{results}, The results of the correlation between the arrival distribution of UHE protons and their source distribution. We summarizes this study in section \\ref{conclusion}. ", "conclusions": "\\label{conclusion} In this paper, we calculated the arrival distribution of UHE protons taking into account energy losses and deflections by EGMF during propagation in intergalactic space in order to investigate the possibility that future observations of UHECRs can unveil the local structure of UHE universe. In order to reproduce a realistic situation, we adopted a structured EGMF model and source distributions which reproduce the observed local structures. The arrival distribution of UHE protons was compared statistically to their source distribution using the correlation coefficients. As the number of observed events increases, the correlation coefficient increases and converges to some value which represents the ability to unveil the source distribution by UHE protons, i.e. charged particles. Thus, the number of events that the correlation coefficient starts to converge is an important number for UHECR observations. In other words, detection of such number of events allows us to unravel UHECR source distribution. We found that UHECRs above $10^{19.8}$eV are best indicators to decipher their source distribution within 100 Mpc from discussion based on the final values of the correlation coefficients and GZK mechanism, and 5000, 500, and 200 event detections above $10^{19.8}$eV on all the sky can unveil their source distribution for the source number densities of $10^{-4}$, $10^{-5}$, and $10^{-6}~{\\rm Mpc}^{-3}$ respectively. Note that ground based detectors observe only about half hemisphere, so only half of such numbers are requested. In this study, we took only EGMF into account as magnetic field, i.e., neglected GMF GMF deflects trajectories of UHE protons efficiently by its regular components, which consist in spiral and dipole components \\citep*{alvarez02,yoshiguchi03b}. A turbulent component of GMF very weakly change the arrival directions of UHE protons \\citep*{yoshiguchi04}. The deflection angles of UHECR protons are a few degree at around $10^{20}$eV except for the direction of Galactic Center. Such deflection disturbs the spatial pattern of UHECR arrival distribution at a few degree scale. The effect of GMF is one of our future investigations. A lot of inquiries on UHECR source number density result in around $10^{-5}~{\\rm Mpc}^{-3}$ on ground of the AGASA results as introduced in section \\ref{intro}. If this number density is true, five year observation by Auger and future observation by TA and Extreme Universe Space Observatory \\citep*{euso} will reveal the distribution of nearby UHECR sources. The dawn of the UHE particle astronomy is just around the corner." }, "0710/0710.1557_arXiv.txt": { "abstract": "\\noindent The field of astroparticle physics is currently developing rapidly, since new experiments challenge our understanding of the investigated processes. Three messengers can be used to extract information on the properties of astrophysical sources: photons, charged Cosmic Rays and neutrinos. This review focuses on high-energy neutrinos ($E_{\\nu}>100$~GeV) with the main topics as follows. \\begin{itemize} \\item The production mechanism of high-energy neutrinos in astrophysical shocks. The connection between the observed photon spectra and charged Cosmic Rays is described and the source properties as they are known from photon observations and from charged Cosmic Rays are presented. \\item High-energy neutrino detection. Current detection methods are described and the status of the next generation neutrino telescopes are reviewed. In particular, water and ice Cherenkov detectors as well as radio measurements in ice and with balloon experiments are presented. In addition, future perspectives for optical, radio and acoustic detection of neutrinos are reviewed. \\item Sources of neutrino emission. The main source classes are reviewed, i.e.~galactic sources, Active Galactic Nuclei, starburst galaxies and Gamma Ray Bursts. The interaction of high energy protons with the cosmic microwave background implies the production of neutrinos, referred to as GZK neutrinos. \\item Implications of neutrino flux limits. Recent limits given by the AMANDA experiment and their implications regarding the physics of the sources are presented. \\end{itemize} ", "introduction": " ", "conclusions": "" }, "0710/0710.2232_arXiv.txt": { "abstract": "We have detected 523 sources in a survey of the Small Magellanic Cloud (SMC) Wing with {\\it Chandra}. By cross-correlating the X-ray data with optical and near-infrared catalogues we have found 300 matches. Using a technique that combines X-ray colours and X-ray to optical flux ratios we have been able to assign preliminary classifications to 265 of the objects. Our identifications include four pulsars, one high-mass X-ray binary (HMXB) candidate, 34 stars and 185 active galactic nuclei (AGNs). In addition, we have classified 32 sources as 'hard' AGNs which are likely absorbed by local gas and dust, and nine 'soft' AGNs whose nature is still unclear. Considering the abundance of HMXBs discovered so far in the Bar of the SMC the number that we have detected in the Wing is low. ", "introduction": "Multi-wavelength studies of the Small Magellanic Cloud (SMC) have shown that it contains a large number of X-ray binary pulsars. From analysis of H$\\alpha$ measurements \\citep{ken91} and supernova birth rates \\citep{fil98} the star formation rate (SFR) for the SMC is estimated to lie in the range 0.04--0.4 $M_{\\odot}$ yr$^{-1}$. \\citet{sht05} used these upper and lower SFR estimates and the linear relation between the number of high-mass X-ray binaries (HMXBs) and the SFR of the host galaxy from \\citet{gri03} to predict the number of HMXBs expected in the SMC with luminosities $\\ge 10^{35}$ erg s$^{-1}$. They found that between 6 and 49 of these systems should be present. Currently $\\sim60$ known or probable HMXBs have been detected in the SMC \\citep[see e.g.][]{hab04,coe05,mcg07}. It is believed that the considerable number of pulsars can be explained in terms of a dramatic phase of star formation, probably related to the most recent closest approach of the SMC and the Large Magellanic Cloud \\citep[LMC;][]{gar96}. To date most of the X-ray studies of the SMC have concentrated on the Bar which has proved to be a significant source of HMXBs. These systems not only provide an homogeneous sample for study, but also give direct insights into the history of our neighbouring galaxy as they are tracers of star formation rates. Part of the puzzle of the X-ray population of the SMC is the missing or under represented components. In particular, there are no known low-mass X-ray binaries (LMXBs) or black hole binaries and only one confirmed supergiant X-ray binary detected to date \\citep[see also][]{mcb07a}. A survey of the X-ray binary population of the LMC by \\citet{neg02} revealed a similar distribution (within small number statistics) to that in our galaxy - all types were present. It is therefore important to try and identify the ``missing'' X-ray binary types in the SMC. We recently completed the first X-ray survey in the SMC Wing with \\chan\\ (see Section \\ref{sect:obs} for more details). A study of the brightest ($>50$ counts) X-ray sources uncovered two new pulsars, and detected two previously known pulsars \\citep{mcg07}. In addition to the four pulsars, the sample included two foreground stars, 12 probable AGNs and five unclassified sources. We found that the pulsars had harder spectra than the other bright X-ray sources. In this paper we report on the analysis of the whole survey and present preliminary classifications for a large fraction of the sources detected. \\begin{figure} \\includegraphics[width=84mm]{fig1.eps} \\caption{The location of the 20 fields studied by Chandra in this work, overlaid on a neutral hydrogen density image of the SMC \\citep{sta99}. The Wing and Bar of the SMC are marked.} \\label{fig:smc_fields} \\end{figure} ", "conclusions": "\\label{sect:disc} For the 523 sources detected in the SMC Wing survey we have been able to find optical matches for 300 of them, and assign preliminary classifications to 265 objects. Our classification method has the advantage that it does not require optical spectra, however, it still requires optical counterparts to be identified. We also note that to classify the remaining 49\\% of the survey deeper optical surveys are needed, and in some cases better coverage of the Wing. The majority of the Wing sources are found to be AGNs. In the whole survey we only identify four pulsars \\citep[see][]{mcg07} and one HMXB candidate, which compared to the Bar is a small sample. The relatively few pulsars detected in the Wing is perhaps not surprising given the accepted link between regions of H$\\alpha$ and star formation, with the main regions of star formation coinciding with the high density H$\\alpha$ region in the Bar \\citep{ken95}. However, in general, the pulsars we detected in the Wing have harder spectra than those in the Bar. It is also remarkable that the only supergiant system so far detected in the SMC, SMC X-1, lies in the Wing. We note that, despite appearances, the SMC is a very three dimensional object. Studies of the Cepheid population by \\citet{lan86} have revealed that the depth of the SMC is up to 10 times its observed width. The two main structures, the Bar and the Wing, could be separated by 10--20 kpc. Could different populations be represented in the two regions? In the case of the HMXBs if we based our response on the X-ray results alone we could perhaps draw the conclusion that the sources in the Wing and Bar are in fact different. However, taking into account the optical spectral analysis in which the optical counterparts for the pulsars were found to be typical of other HMXBs in the SMC \\citep{sch07,mcb07b}, different populations seem less likely. This could imply that there is absorption local to the sources which effects the X-ray spectral results. There is also the possibility that a greater population of HMXBs does exist in the Wing of the SMC, but we were not fortunate enough to catch more than a handful of them when they were switched on. From our studies of 10 years of {\\it RXTE} data we find that the probability of a Be X-ray transient being in an active phase is only, on average, $\\sim 10$\\% \\citep[Figure 4.62,][]{gal06}. Quiescent X-ray transients have been detected previously in the Milky Way with luminosities $<10^{34}$ erg s$^{-1}$ \\citep[e.g.][]{neg00,cam02}. The origin of the quiescent luminosity in Be X-ray transients is still under debate, with a number of processes suggested to account for the detected emission \\citep[see e.g.][]{cam02,kre04}. The two mechanisms detectable from sources located in the SMC are: accretion onto the magnetospheric boundary, the propeller regime \\citep{ill75,cam00}, and very low rate accretion onto the surface of the neutron star, i.e. residual/leaking accretion \\citep[e.g.][]{ste94}. The one HMXB candidate that we have identified has a luminosity (at the distance to the SMC) of $3.2\\times10^{33}$ erg s$^{-1}$ so it could be a quiescent source. The lack of HMXBs in the Wing indicates that we are looking at an older population which is confirmed by optical studies of the star formation history of the SMC \\citep[e.g.][]{har04}. In theory this should increase our chances of detecting LMXBs. Arguably, LMXBs should be well distributed within the SMC, i.e. they should lie in the Bar and the Wing, however, deep looks of the SMC Bar \\citep{naz03} have been unsuccessful in detecting any. The number of LMXBs expected in the SMC is proportional to the total stellar mass of the galaxy, resulting in a prediction of only one system with an X-ray luminosity of $\\geq 10^{35}$ erg s$^{-1}$ \\citep[see][]{sht05}. However, \\citet{gar01} have shown that quiescent LMXBs can be as faint as $2\\times 10^{30}$ erg s$^{-1}$. To go as deep as that is beyond the capability of current X-ray telescopes, but in 100 ks it would be possible to reach a limit of $\\sim 10^{32}$ erg s$^{-1}$, sufficient to detect a sample of fainter sources and study their characteristics. If an observation like this were performed in the Wing it could be compared directly with the deep exposures of the Bar \\citep{naz03,zez05} and help quantify the LMXB population in the SMC." }, "0710/0710.0371_arXiv.txt": { "abstract": "One well-known way to constrain the hydrogen neutral fraction, $\\bxhi$, of the high-redshift intergalactic medium (IGM) is through the shape of the red damping wing of the \\lya absorption line. We examine this method's effectiveness in light of recent models showing that the IGM neutral fraction is highly inhomogeneous on large scales during reionization. Using both analytic models and ``semi-numeric\" simulations, we show that the ``picket-fence\" absorption typical in reionization models introduces both scatter and a systematic bias to the measurement of $\\bxhi$. In particular, we show that simple fits to the damping wing tend to \\emph{overestimate} the true neutral fraction in a partially ionized universe, with a fractional error of $\\sim 30\\%$ near the middle of reionization. This bias is generic to any inhomogeneous model. However, the bias is reduced and can even underestimate $\\bxhi$ if the observational sample only probes a subset of the entire halo population, such as quasars with large HII regions. We also find that the damping wing absorption profile is generally steeper than one would naively expect in a homogeneously ionized universe. The profile steepens and the sightline-to-sightline scatter increases as reionization progresses. Of course, the bias and scatter also depend on $\\bxhi$ and so can, at least in principle, be used to constrain it. Damping wing constraints \\emph{must} therefore be interpreted by comparison to theoretical models of inhomogeneous reionization. ", "introduction": "\\label{intro} The reionization of hydrogen in the intergalactic medium (IGM) is a landmark event in the early history of structure formation, because it defines the moment at which galaxies (and black holes) affected every baryon in the Universe. As such, it has received a great deal of attention -- both observationally and theoretically -- in the past several years. Unfortunately, the existing observational evidence is enigmatic (see \\citealt{fan06-review} for a recent review). Electron scattering of cosmic microwave background photons implies that reionization occurred at $z \\sim 10$, albeit with a large uncertainty \\citep{page06}. On the other hand, \\lya forest spectra of quasars at $z \\sim 6$ show some evidence for a rapid transition in the globally-averaged neutral fraction, $\\bxhi$ (e.g., \\citealt{fan06}). However the \\lya absorption is so saturated in the \\citet{gunn65} trough (with optical depth $\\tau_{\\rm GP} \\ga 10^5 \\bxhi$) that constraints derived from that spectral region \\citep{fan06, maselli07} are difficult to interpret (e.g, \\citealt{lidz06, becker07}). Another probe is the red damping wing of the IGM \\lya absorption: the line is so saturated at these redshifts that even photons that are emitted redward of the \\lya resonance can suffer significant absorption from the strong damping wings of that transition. This has a number of consequences for high-redshift observations. For example, surveys that search for high-$z$ galaxies through their \\lya emission lines will find fewer and fewer galaxies as the IGM becomes more and more neutral \\citep{haiman02-lya, santos04}, although galaxy clustering strongly moderates this decline \\citep{furl04-lya, furl06-lya, mcquinn07, mesinger07-lya}. Such surveys have now detected objects at $z \\sim 6.5$--$9$ (e.g., \\citealt{kashikawa06, iye06, stark07}), but their implications for reionization are unclear \\citep{malhotra04, haiman05-lya, malhotra06, kashikawa06, dawson07, dijkstra07, mcquinn07-lya, mesinger07-lya}. The evolution of galaxy abundances and clustering measures the damping wing absorption in a statistical sense, but even more information can potentially be gleaned from the damping wing absorption profiles in individual objects \\citep{miralda98}. For the galaxies described above, this information is difficult to extract because of their faintness and the complicated origins of their \\lya emission lines \\citep{mcquinn07-lya}. However, high signal-to-noise spectra of bright objects could be extremely helpful. If the damping wing profile from IGM absorption can be isolated from these spectra, this would provide detailed information on the neutral gas along each particular line of sight (LOS) -- rather than the statistical information available from most other probes. This is very useful, as reionization is expected to be highly inhomogeneous. There are two candidates for such high signal-to-noise spectra at high-redshifts: quasars and gamma-ray bursts (GRBs). Quasars present several challenges: complicated intrinsic spectra, biased IGM environments \\citep{barkana04-grb, lidz07}, and large HII regions (which significantly weaken the damping wing absorption redward of the quasar \\lya line, and can necessitate detailed spectral analysis of the blue side of the line; \\citealt{madau00, mesinger04-mockprox}). Nevertheless, there have already been two claims of damping wing detections in high-redshift spectra, both using quasars from the Sloan Digital Sky Survey (SDSS). \\citet{mesinger04} detected a $\\bxhi \\gsim 0.2$ damping wing through the decreased fluctuations in the total Ly$\\alpha$ optical depth near the edge of the HII region surrounding \\qnametwoeight\\ ($z_S=6.28$). Similarly, by simulating the optical depth distributions blueward of the Ly$\\alpha$ line center and comparing them with deep observations, \\citet{mesinger07-prox} detected the presence of a $\\bxhi \\gsim 0.033$ damping wing in the spectra of \\qnametwoeight\\ and \\qnametwotwo\\ ($z_S=6.22$). The maximum likelihood was at $\\bxhi=1$ for both quasars. The second set of candidates, GRBs, have fewer obstacles to overcome. Long-duration GRBs are believed to be remnants of massive stars (and so trace the bulk of the star formation, which probably occurs in lower-mass halos with more ``typical\" IGM environments), and their afterglows have extremely simple power-law intrinsic spectra (see, e.g., \\citealt{piran05} for a review). The event rates at high redshifts may be quite high, and cosmological time-dilation helps to identify the sources when they are still bright \\citep{bromm02-grb, ciardi00, lamb00, mesinger05}. As a result, there is a great deal of optimism in the literature regarding their potential for damping-wing measurements (e.g. \\citealt{miralda98, barkana04-grb}). The highest-redshift GRB afterglow observed so far (at $z \\approx 6.3$), has already been used to constrain the global neutral fraction at that time \\citep{kawai06, totani06}. Unfortunately, this object illustrates the major difficulty with the red damping wing test for GRBs: intrinsic absorption in the host galaxies \\citep{miralda98}. Most GRBs are now known to have large columns of associated neutral hydrogen \\citep{vreeswijk04, chen04-grb}. Roughly $20\\%$ of well-studied objects have $N_{\\rm HI} \\la 10^{20} \\colden$ \\citep{chen07}, although nearly all of the objects in this sample are at $z \\la 6$. The $z \\approx 6.3$ GRB does appear to have intrinsic absorption with $N_{\\rm HI} \\sim 10^{21.6} \\colden$ \\citep{totani06}, which makes it difficult to constrain the IGM absorption. In principle, it is still possible because isolated HI absorbers have different spectral profiles than the IGM (with the optical depth inversely proportional to the wavelength offset squared for isolated absorbers, and to the wavelength offset itself for the IGM). The two sources can then be separated by looking at the shape of the absorption. \\citet{totani06} found a best fit with $\\bxhi=0$ and estimated that $\\bxhi \\la 0.17$ ($0.60$) at 68\\% (95\\%) confidence. Better constraints will require faster followup (when the afterglow is brighter) and systems with less intrinsic absorption. To date, the red damping wing test has generally been assumed to be simple and straightforward. It is usually argued that the absorption is sensitive to a large path length in the IGM, so that small-scale clumpiness can be ignored and that the ionized fraction can be taken to be uniform (for an exception, see \\citealt{barkana02}). However, most models of reionization have much more inhomogeneous distributions of neutral and ionized gas, with discrete HII regions surrounding clusters of galaxies, and a sea of nearly neutral gas separating them (e.g., \\citealt{arons72, shapiro87}). Such a picture is inevitable when hot stars ionize the gas. Moreover, the most recent models show that the ionized bubbles can become quite large even relatively early in reionization, with sizes $\\ga 10 \\Mpc$ when $\\bxhi \\sim 0.5$ \\citep{furl04-bub, furl05-charsize, iliev05-sim, zahn07-comp, mcquinn07, mesinger07}. Because the damping wing is sensitive to fluctuations on Mpc scales, it is actually not a good approximation to take the IGM ionized fraction to be constant. In this paper, we will examine whether (and how) the damping wing can actually be used to constrain the reionization process. We summarize the basic physics of the line in \\S \\ref{lya}. We then examine a series of toy models of the ``picket-fence\" absorption typical of the IGM during reionization in \\S \\ref{toy}. In particular, we show that interpreting measurements with the naive view of a uniform IGM is not only subject to significant scatter (from the different networks of ionized bubbles intersected along different lines of sight) but also a substantial systematic bias. In \\S \\ref{sim}, we describe the ``semi-numeric\" simulations used to generate our main results, which we present in \\S \\ref{results}. This more detailed picture confirms that scatter between different lines of sight and bias relative to the naive view will be critical in interpreting any observed sources. Finally, we conclude in \\S \\ref{disc}. When this project was nearing completion, we learned of a similar effort by \\citet{mcquinn07-damp} and refer the reader there for a complementary discussion. In our numerical calculations, we assume a cosmology with $\\Omega_m=0.26$, $\\Omega_\\Lambda=0.74$, $\\Omega_b=0.044$, $H=100 h \\hunits$ (with $h=0.74$), $n=0.95$, and $\\sigma_8=0.8$, consistent with the most recent measurements \\citep{spergel06}. Unless otherwise specified, we use comoving units for all distances. ", "conclusions": "\\label{disc} In this paper, we have examined how the shape of the \\lya red damping wing can be used to constrain the IGM before reionization is complete. In the past, it has usually been assumed that the absorbing gas can be well-approximated by a uniform density medium with constant ionized fraction. However, recent reionization models have shown that ionized bubbles can be quite large, so the latter is not a good approximation. We have therefore critically examined how well the damping wing constrains the neutral fraction during inhomogeneous reionization. We have identified two major issues with its interpretation. First, there is substantial scatter in the optical depth along different lines of sight. Most of this is due to the scatter in the distance between the source and the nearest patch of neutral gas; however, there is still non-negligible scatter even if this distance can be measured from the shape of the damping wing. In our semi-numeric simulations, the fractional r.m.s. fluctuation in $\\bxhi$ thus estimated increases from 0.1 to 1 over the range $0.9\\gsim\\bxhi\\gsim0.2$. Fortunately, this statistical uncertainty can be reduced simply by finding more lines of sight. The other problem is more severe: we have shown that the ``picket-fence\" absorption from inhomogeneous reionization adds a systematic, and often large, \\emph{bias} to measurements of the neutral fraction. Although the damping wing is indeed sensitive to a large path length through the IGM, it is most sensitive to the closest gas. As a result, simple fits to the damping wing will always \\emph{overestimate} the true neutral fraction in a partially ionized universe, with an error of $\\sim 30\\%$ near the middle of reionization. This bias is generic to any inhomogeneous model. The bias is reduced and can even become negative if observations only probe a subset of the entire halo population, such as quasars with large HII regions. Both the systematic and statistical uncertainty can be reduced by a careful fit to the damping wing spectral profile, which is typically steeper than the naively expected $(\\Delta \\lobs)^{-1}$ profile. However, because the absorption typically comes from many neutral patches, a large number of parameters are required for a detailed fit, and given the relatively modest difference from the expected behavior, these will be difficult to measure, probably only possible in systems with intrinsically large optical depths. Moreover, the scatter in the profiles, even at fixed $\\tau_D$, is sufficient that large samples will be required to put strong constraints on reionization from the spectral shape. Of course, the bias and scatter also depend on $\\bxhi$ and so can, at least in principle, be used to constrain it. For example, large dispersion in the inferred neutral fractions could be an indicator of $\\bxhi \\la 0.2$. If an independent estimate of $\\bxhi$ exists, one could reverse the direction of analysis, and use the bias and scatter to constrain the reionization model and topology. Fortunately, for a given model of reionization, the dispersion and bias can be calibrated by theoretical models. We therefore argue that the most efficient way to constrain reionization with the damping wing is through comparison with detailed models. Of course, any such constraints will be model-dependent, but we believe that the morphology of reionization is now sufficiently well-understood (see, e.g., \\citealt{furl05-charsize, mcquinn07}) that these uncertainties will likely not dominate the statistical uncertainties from the small number of accessible sources, at least in the relatively near future. For example, the reionization morphology is nearly independent of redshift \\citep{furl04-bub, mcquinn07}. Also, we have found only a modest dependence of the $x_D$ distribution on halo mass (mostly due to the variation in bubble size with mass). However, toward the end of reionization, when the absorption is dominated by rare, narrow sheets of neutral hydrogen, the details of the radiative transfer algorithm (or an approximation to it, as in our models) and of the sample selection will be extremely important. Nevertheless, the task is challenging, as the damping wing profile must be separated from the rapidly varying resonance absorption for quasars (as in \\citealt{mesinger04, mesinger07-prox}) or from intrinsic absorbers for GRBs. Fortunately, in the latter case $\\sim 20\\%$ of moderate-redshift GRBs have only modest absorbers and will still be useful for these purposes \\citep{chen07}. So far, the damping wing analysis has been performed on three high-redshift quasars: \\qnamefourtwo\\ ($z_S=6.42$), \\qnametwoeight\\ ($z_S=6.28$), \\qnametwotwo\\ ($z_S=6.22$) \\citep{mesinger04, mesinger07-prox}, as well as GRB 050904 ($z_S \\approx 6.3$) \\citep{kawai06, totani06}. This paper highlights the need to calibrate these and future damping wing analysis with simulations of the reionization morphology. Obviously we cannot set firm constraints without detailed simulations of the observations. Nevertheless, the mean bias we find from our simulations seems to work in the direction of strengthening the upper limit (on $\\bxhi$) from the \\citet{totani06} measurements, and weakening the lower limit from the \\citet{mesinger04, mesinger07-prox} constraints at $\\bxhi\\lsim0.6$ (although, interestingly, it would strengthen them if $\\bxhi\\gsim0.6$; see the sign change for the bias in Fig.~\\ref{fig:high_mass_mean_sig}). Conversely, the steeper-than-expected absorption profile seems to work in the direction of weakening the \\citet{totani06} constraints (especially because it must be distinguished from strong internal absorption) while strengthening the \\citet{mesinger04, mesinger07-prox} constraints. The absorption profile might be more relevant than the bias for these studies, as an overall bias can be partially degenerate with other free parameters in the fit: that is, when the absorption profile \\emph{can} be detected to high precision, its shape will certainly be useful in constraining $\\bxhi$. The scatter in both effects would probably somewhat erode the confidence contours for all of these studies. On the other hand, our model predicts large scatter between different LOSs at the end of reionization, which is consistent with the measurements at $z \\sim 6.3$. More precise limits will require simultaneous fits to the intrinsic absorption and the range of possible IGM absorber profiles, and we defer them to future work. Another intriguing possibility is to try to measure damping wing characteristics from stacked spectra of many \\lyans-emitting galaxies. \\citet{mcquinn07} have shown that the wing shape is difficult to separate from uncertainties in the line for individual objects, and the scatter we have described will also make the interpretation of individual faint emitters problematic. But, if the characteristics of the population are relatively constant, stacking may increase the signal to noise sufficiently to allow a detection of a ``mean\" damping wing at each redshift, even far redward of line center. SRF thanks Crystal Martin and Josh Bloom for conversations that stimulated this work. We thank Z. Haiman for helpful comments on this manuscript. This research was partially supported by grant NSF-AST-0607470." }, "0710/0710.2697_arXiv.txt": { "abstract": "We studied the clustering properties and multiwavelength spectral energy distributions of a complete sample of 162 Ly$\\alpha$-emitting (LAE) galaxies at $z\\simeq 3.1$ discovered in deep narrow-band MUSYC imaging of the Extended Chandra Deep Field South. LAEs were selected to have observed frame equivalent widths $>$80\\AA\\ and emission line fluxes $>1.5\\times10^{-17}$ergs cm$^{-2}$ s$^{-1}$. Only 1\\% of our LAE sample appears to host AGN. The LAEs exhibit a moderate spatial correlation length of $r_0=3.6^{+0.8}_{-1.0}$Mpc, corresponding to a bias factor $b=1.7^{+0.3}_{-0.4}$, which implies median dark matter halo masses of $\\log_{10}\\mathrm{M_{med}} = 10.9^{+0.5}_{-0.9}$M$_\\odot$. Comparing the number density of LAEs, $1.5\\pm0.3\\times10^{-3}$Mpc$^{-3}$, with the number density of these halos finds a mean halo occupation $\\sim$1--10\\%. The evolution of galaxy bias with redshift implies that most $z=3.1$ LAEs evolve into present-day galaxies with $L<2.5L^*$, whereas other $z>3$ galaxy populations typically evolve into more massive galaxies. Halo merger trees show that $z=0$ descendants occupy halos with a wide range of masses, with a median descendant mass close to that of $L^*$. Only 30\\% of LAEs have sufficient stellar mass ($>\\sim3\\times10^9$M$_\\odot$) to yield detections in deep Spitzer-IRAC imaging. A two-population SED fit to the stacked $UBVRIzJK$+[3.6,4.5,5.6,8.0]$\\mu$m fluxes of the IRAC-undetected objects finds that the typical LAE has low stellar mass ($1.0^{+0.6}_{-0.4}\\times10^9$M$_\\odot$), moderate star formation rate ($2\\pm1 $M$_\\odot$yr$^{-1}$), a young component age of $20^{+30}_{-10}$Myr, and little dust ($A_V<0.2$). The best fit model has 20\\% of the mass in the young stellar component, but models without evolved stars are also allowed. ", "introduction": "\\label{sec:intro} The discovery of high-redshift Ly$\\alpha$-emitting galaxies (LAEs) opened a new frontier in astronomy \\citep{cowieh98,huetal98}. Because the Ly$\\alpha$ line is easily quenched, a galaxy with detectable Ly$\\alpha$ emission is likely dust-free, i.e., in the initial phases of a burst of star formation. The Ly$\\alpha$ lines have large equivalent widths ($20${\\AA}$<$EW$_{\\mathrm{rest}}<\\sim 100${\\AA}) and broad velocity widths (100 km~s$^{-1}$$<$FWHM$<$800 km~s$^{-1}$) and are often asymmetric, indicative of high-redshift galaxies undergoing active star-formation \\citep[e.g.,][]{manningetal00, kudritzkietal00,arnaboldietal02,rhoadsetal03, dawsonetal04,huetal04,venemansetal05,matsudaetal06,gronwalletal07}. Other high-redshift galaxy populations (including Lyman break galaxies, Distant Red Galaxies, Sub-Millimeter Galaxies) exhibit strong clustering and should evolve into elliptical and giant elliptical galaxies today \\citep{adelbergeretal05a,quadrietal07a,webbetal03}. These objects were selected by unusually strong rest-frame continuum emission in the ultraviolet, optical, and far-IR respectively, resulting in $10^{10}L_\\odot \\leq L_{bol} \\leq 10^{12}L_\\odot$ \\citep{reddyetal06b}. Such strong continua appear to occur primarily in deep potential wells that are strongly biased versus the general distribution of dark matter halos. LAEs instead offer the chance to probe the faint end of the (bolometric) luminosity function of high-redshift galaxies, which contains the majority of galaxies. The strong Ly$\\alpha$ emission line allows detection and spectroscopic confirmation of LAEs with typical bolometric luminosities $\\simeq 10^{10} L_\\odot$. A detailed calculation of the LAE luminosity function at $z\\simeq 3.1$ is given in \\citet{gronwalletal07}. Spectral energy distribution (SED) modelling of the stacked $UBVRIzJK$ photometry of 18 LAEs in the Extended Chandra Deep Field-South (ECDF-S) \\citep{gawiseretal06b} showed the average galaxy to have low stellar mass ($<10^9$ M$_\\odot$) and minimal dust \\citep[see also][]{nilssonetal07}. LAEs have the highest {\\it specific} star formation rates (defined as SFR divided by stellar mass) of any type of galaxy, implying the youngest ages \\citep{castroceronetal06}. Because Ly$\\alpha$ emission is easily quenched by dust, LAEs have often been characterized as protogalaxies experiencing their first burst of star formation \\citep[e.g.][]{hum96}. However, the differing behavior of Ly$\\alpha$ and continuum photons encountering dust and neutral gas makes it possible for older galaxies to exhibit strong Ly$\\alpha$ emission when morphology and kinematics favor the escape of these photons towards Earth \\citep{neufeld91,haimans99,hanseno06}. This could allow older, dusty galaxies with actively star-forming regions to exhibit Ly$\\alpha$ emission with high equivalent width. SED modelling of LAEs using Spitzer-IRAC \\citep{fazioetal04} to probe rest-frame near-infrared wavelengths, where old stars dominate the emission, has yielded mixed results. \\citet{pirzkaletal07} report extremely young ages of a few Myr and low stellar masses ($10^6$M$_\\odot<$M$_*<10^8$M$_\\odot$) from SED modelling of 9 LAEs at $4.010^{10.6}$M$_\\odot$ at $z=3.1$ (defined as the age since half of the dark matter mass was accumulated) to be $\\sim$600 Myr, with only $<$10\\% of halos younger than 100 Myr. If repeated LAE phases occur, the mean halo occupation of $\\sim1$--$10$\\% can be interpreted as a ``duty cycle'' telling us what fraction of each halo's lifetime is spent in the early phases of starbursts before significant dust is generated, and the population-averaged young age of $\\sim$20 Myr would imply that this phase typically lasts $\\sim$40 Myr. Alternatively, if all dark matter halos experience a single LAE phase shortly after their ``formation'' in a major merger, the mean halo occupation implies that LAEs will only be found in the youngest $1$--$10$\\% of halos, which is barely consistent with their single-population best-fit age of 100 Myr. If LAEs represent a subset of dark matter halos selected to have ages less than 100 Myr, their observed clustering may underestimate their dark matter halo masses by up to a factor of two \\citep[see][]{gaow07}. \\begin{figure}[h!] \\epsscale{1.25} \\plotone{f6.eps} \\caption[] { Tracks show the evolution of bias with redshift calculated using the no-merging model. The filled circle shows our result for the bias of LAEs at $z=3.1$. Previous results at high-redshift are shown for LAEs at $z=4.5$ and $z=4.86$ (stars, \\citealp{kovacetal07} and \\citealp{ouchietal03}, respectively), LBGs at $z\\sim4$ (stars, \\citealp{ouchietal04b}), K-selected galaxies (diamonds, \\citealp{quadrietal07a}), bright LBGs at $z\\sim 3$ (triangle, \\citealp{leeetal06}), and BM, BX, and LBG galaxies (asterisks, \\citealp{adelbergeretal05b}, \\citealp{adelbergeretal05a}). Local galaxy clustering is shown for SDSS galaxies (open circles, \\citealp{zehavietal05}) and for rich clusters (cross, \\citealp{bahcalletal03}). K-band limits are in Vega magnitudes. } \\label{fig:bias} \\end{figure} Figure \\ref{fig:bias} shows the reported bias values for LAEs to be lower than those of other $z>3$ galaxy populations \\citep[bias values determined as in][]{quadrietal07a}. The expected evolution of bias is shown for the ``no-merging'' model \\citep{fry96,whiteetal07}. A realistic amount of merging will cause the bias to drop somewhat faster, so the plotted trajectories provide an upper limit on the bias factor of a given point at lower redshifts. This shows that typical $z=3.1$ LAEs will evolve into galaxies of at most a few times $L^*$ at $z=0$. The bias values imply that LAEs at $z=3.1$ might evolve into the subset of BX galaxies at $z\\simeq 2.2$ dimmer than $K=21.5$, which also show relatively weak clustering \\citep{adelbergeretal05a}. The $K>21.5$ BX galaxies have average M$_*=1.5\\times10^{10}$M$_\\odot$, so the $z=3.1$ LAEs would need to form stars at an average rate of 14 M$_\\odot$yr$^{-1}$ over the intervening Gyr. This could be achieved with a constant specific SFR and no merging or with a constant SFR and $\\sim$2 major mergers. The only previous measurements of LAE clustering in unbiased fields are at $z=4.5$ \\citep{kovacetal07} and $z=4.86$ \\citep{ouchietal03}, and this earlier LAE population appears to have significantly stronger clustering, consistent with possibly evolving into typical Lyman break galaxies at $z\\simeq 3$. The models of \\citet{ledelliouetal06} predict stellar and dark matter masses and star formation rates for $z=3$ LAEs within a factor of two of our results, despite assuming a top-heavy IMF and a very low escape fraction $f_{esc}=0.02$ that appears inconsistent with the observed lack of dust \\citep[see][for an alternative approach]{kobayashietal07}. \\citet{maoetal07} used the stellar mass of $5\\times10^8$M$_\\odot$ observed by \\citet{gawiseretal06b} to predict LAE dark matter masses of $10^{10}<$M$<10^{11}$M$_\\odot$, in the lower end of our allowed range. The stellar ages of $\\sim 20$ Myr preferred by the two-population fit are noticeably lower than the maximum values of 100 to 500 Myr predicted by these authors, \\citet{moriu06}, and \\citet{haimans99}, but the ages of 60 to 350 Myr preferred for the case of no evolved stars would be compatible. None of the current models and numerical simulations of LAEs \\citep[see also][]{thommesm05,razoumovs06,tasitsiomi06} predict their present-day descendants. Nonetheless, the evolution of a significant fraction of $z=3.1$ LAEs into $z=0$ $L^*$ galaxies with dark matter mass $M_{DM}\\simeq2\\times10^{12}$M$_\\odot$ and stellar mass $M_*\\simeq4\\times10^{10}$M$_\\odot$ \\citep{ichikawaetal07} appears reasonable. Fig. \\ref{fig:mass} shows the histogram of present-day masses of dark matter halos in the Milli-Millenium merger trees that have progenitors with M$>5\\times10^{10}$M$_\\odot$ at $z=3.1$. The median present-day halo mass is $1.2\\times10^{12}$M$_\\odot$, and this would increase if LAEs found in sub-halos of massive $z=3.1$ halos were included. \\citet{lietal07} predict that the main progenitor of a present-day $L^*$ galaxy had a dark matter mass of $\\sim10^{11}$M$_\\odot$ at $z=3$ and that these galaxies experienced several major mergers at $1.55\\times10^{10}$M$_\\odot$ at $z=3.1$. The dashed lines show the median halo mass of $1.2\\times10^{12}$M$_\\odot$ and the 25th and 75th percentile values of $2.9\\times10^{11}$M$_\\odot$ and $7.6\\times10^{12}$M$_\\odot$. } \\label{fig:mass} \\end{figure} The typical LAE stellar mass at $z=3.1$ is lower than that of any other studied high-redshift population \\citep[see][]{reddyetal06b} but is close to that of dim ($i<26.3$) Lyman break galaxies (LBGs) at $z\\sim5$ \\citep{vermaetal07}. LAEs at $z=3.1$ have much lower star formation rate, stellar age, stellar mass, dark matter halo mass, and dust extinction than the $\\sim30$M$_\\odot$yr$^{-1}$, $\\sim500$ Myr, $\\sim2\\times10^{10}$M$_\\odot$, $\\sim3\\times10^{11}$M$_\\odot$, $A_V\\simeq1$ LBG population at $z\\sim3$ \\citep[$R<25.5$,][]{shapleyetal01,adelbergeretal05a} or the $\\sim100$M$_\\odot$yr$^{-1}$, $\\sim2$ Gyr, $\\sim10^{11}$M$_\\odot$, $\\sim10^{13}$M$_\\odot$, $A_V\\simeq2.5$ Distant Red Galaxy (DRG) population \\citep{webbetal06,forsterschreiberetal04,quadrietal07a}. The high-redshift Sub-Millimeter Galaxies \\citep{chapmanetal03b} appear to be the most massive and dusty, with the highest SFR. LAEs may represent the beginning of an evolutionary sequence where galaxies gradually become more massive and dusty due to mergers and star formation, but most LAEs at $z=3.1$ will never reach the DRG stage since DRG stellar and dark matter masses are already greater than those of present-day $L^*$ galaxies. The Damped Ly$\\alpha$ Absorption systems (DLAs, \\citealp{wolfeetal05}) are another high-redshift population that probes the faint end of the luminosity function. The dark matter halo masses of DLAs at $z\\sim3$ were determined by \\citet{cookeetal06b} to lie in the range $10^9<$M$<10^{12}$M$_\\odot$ i.e., $1.3$0.5 \\citep{abraham96, vanden96, abraham99}. These authors concluded that at lookback times greater than 5 Gyr disks were either dark matter dominated or dynamically too hot (perhaps due to the increased merging activity) to host bars. However, the small volume probed by the HDFs (only thirty bright, face-on spiral galaxies between $0$0.7. Overall, the fraction of barred spirals in the NICMOS HDF remained extremely low, as in the optical HDF studies. But \\citet{sheth03} noted that their study was limited by the coarse NICMOS resolution (0$\\farcs$15) such that only the largest (and rarest) bars could be identified (bars with semi-major axis $>$5 kpc). When the fraction of these large bars at z$>$0.7 was compared to local samples, there was no compelling evidence for a decline in barred spirals, but likewise the NICMOS data did not unveil any new bars at low redshifts; all except one of the four bars in the \\citet{sheth03} study are at z$>$0.9, where k-correction effects are important (\\S \\ref{bandshifting}). A major advance in spatial resolution was possible with the Advanced Camera for Surveys (ACS) whose 0$\\farcs$05 pixels are able to resolve all but the smallest (nuclear, $<$2 kpc diameter) bars at all redshifts. Using ACS data, two studies \\citep{elm04,jogee04} found that contrary to the previous HDF results, the bar fraction is constant at 30\\% over the last 8 Gyr (since $z=1.2$). The sample sizes, however remained modest in these studies (186 in \\citealt{elm04}, and 258 in \\citealt{jogee04}). In this paper we examine in detail the redshift evolution of the bar fraction using the unparalleled wide and deep 2-square degree COSMOS data set. The plan for the paper is as follows: in Section 2 we describe our sample selection procedure. The classification methodology we have adopted is described in Section 3. Our main results are presented in Section 4, before being discussed in Section 5. Our conclusions are summarized in Section 6. An Appendix to this paper provides a detailed analysis of possible selection effects at high redshift and a discussion of our local calibration sample of 139 galaxies from the Sloan Digital Sky Survey (SDSS) Data Release 4 \\citep{adelman06}. Throughout this paper we adopt a flat $\\Lambda$-dominated cosmology with $H_0$=70 km\\ s$^{-1}$\\ Mpc$^{-1}$, $\\Omega_M=0.3$, and $\\Omega_\\Lambda=0.7$. ", "conclusions": "Bars are an important signpost of galaxy evolution because once a galaxy disk is sufficiently massive, dynamically cold and rotationally supported it forms a bar. Therefore the evolution of the bar fraction over time is an important indicator of the evolutionary history of disk galaxies and the assembly of the Hubble sequence. Using a detailed analysis of 2,157 L$^*$ face-on, spiral galaxies from 0.0$<$z$<$0.84 in the COSMOS 2-square degree survey we have investigated the evolution of the bar fraction over the last 7 Gyr. We have undertaken an extensive and careful analysis of selection effects (k-correction, surface brightness dimming, inclination, spatial resolution, etc.) which is detailed in the Appendix. Our main results are as follows: $\\bullet$ The bar fraction for L$^*$ galaxies drops from about 65\\% in the local Universe to about 20\\% at z=0.84. Over this redshift range the fraction of {\\em strong} bars (SB) drops from about 30\\% to under 10\\%. Thus at a lookback time of 7 Gyr, when the Universe was half its present age, fundamental aspects of Hubble's `tuning fork' classification sequence had not yet fallen into place. Only about one fifth of all spiral galaxies were ``mature'' enough (dynamically cold, massive and rotationally supported) to host galactic structures of the type we see today. $\\bullet$ For the total f$_{bar}$ (SB+SAB), the change is far less dramatic between z=0.3 and z=0.0 indicating slow evolution in galactic structures in L$^*$ galaxies over the last 4 Gyr. It is likely that there is significant evolution in the formation of bars in the sub-L$^*$ galaxies over this period. $\\bullet$ One of the most significant findings in this study is the correlation between f$_{bar}$ and the galaxy mass, luminosity and color. We find that in the highest redshift bins f$_{bar}$ is higher in the more massive, luminous and redder systems. In fact, in the most massive systems, f$_{bar}$ is already as high at z=0.8 as the local value. These systems thus had already arrived with their present Hubble types at a lookback time of 7 Gyr. In the subsequent 3 Gyr, from z=0.84 to z=0.3, the lower mass, bluer systems evolved more slowly toward their present Hubble types. Thus the signature of downsizing is intimately connected with dynamical maturity of disks and is present in the formation of galactic structure. $\\bullet$ Finally, we find a slight preference for barred galaxies to be more bulge-dominated in the high redshift bin. This correlation is consistent with the dynamical downsizing found for bars in general if bars and bulges both form earlier and more prominently in the most massive galaxies. The lack of a stronger correlation may be related to the variety of bulges: bars are also likely to be involved with the inflow that builds pseudo-bulges. Given the strong correlation between bulge properties and black hole mass seen today, there may be a co-evolution of bars, bulges and black holes in some galaxies. The exact details of these processes remain to be investigated." }, "0710/0710.3528_arXiv.txt": { "abstract": "{The High Energy Stereoscopic System (H.E.S.S.), located in the Khomas Highlands of Namibia, is an array of four imaging atmospheric-Cherenkov telescopes designed to detect $\\gamma$-rays in the very high energy (VHE; $>$ 100 GeV) domain. Its high sensitivity and large field-of-view (5$^{\\mbox{\\tiny o}}$) make it an ideal instrument to perform a survey within the Galactic plane for new VHE sources. Previous observations in 2004/2005 resulted in numerous detections of VHE gamma-ray emitters in the region l = 330$^{\\mbox{\\tiny o}}$ - 30$^{\\mbox{\\tiny o}}$ Galactic longitude. Recently the survey was extended, covering the regions l = 280$^{\\mbox{\\tiny o}}$ - 330$^{\\mbox{\\tiny o}}$ and l = 30$^{\\mbox{\\tiny o}}$ - 60$^{\\mbox{\\tiny o}}$, leading to the discovery of several previously unknown sources with high statistical significance. The current status of the survey will be presented.} \\begin{document} ", "introduction": "The majority of the newly discovered sources of very high energy (VHE; $>$ 100 GeV) $\\gamma$-rays are related to late phases of stellar evolution, either directly to massive stars or to the compact objects they form after their collapse. The possible associations include pulsar wind nebulae (PWN) of high spin-down luminosity pulsars such as G\\,18.0$-$0.7 \\cite{hess_j1825}, supernova remnants like RX\\,J1713.7$-$3946 \\cite{RXJ1713}, and open star clusters like Westerlund\\,2 \\cite{westerlund2}. As these objects cluster closely along the Galactic plane, a survey of this region is an effective approach to discover new sources and source classes of VHE $\\gamma$-ray emission. ", "conclusions": "The H.E.S.S. Galactic plane survey, which started in the year 2004, now reaches from $-$85$^{\\circ}$ longitude to 60$^{\\circ}$ longitude, and covers an approximately 6$^{\\circ}$ broad band around latitude b = 0$^{\\circ}$. In total, more than 950\\,hours of data were taken in this region, including survey mode observations, re-observations of source candidates and dedicated observations of known or suspected $\\gamma$-ray sources. The first stage of the survey, covering the inner 60$^{\\circ}$ of the Galactic plane, has increased the number of known VHE $\\gamma$-sources within this region from three at the beginning of 2004 to seventeen. Further follow-up observations within this region and the extension of the survey along the Galactic plane resulted in the discovery of even more additional VHE $\\gamma$-ray emitters. Most of them were presented during this conference. Multi-wavelength follow-up observations and archival searches have already begun, and will be crucial for understanding the underlying processes at work in these astrophysical objects." }, "0710/0710.1836_arXiv.txt": { "abstract": "We report on the first detection of maser emission in the $J$=11-10, $J$=14-13 and $J$=15-14 transitions of the $v$=0 vibrational state of SiS toward the C-rich star IRC+10216. These masers seem to be produced in the very inhomogeneous region between the star and the inner dust formation zone, placed at $\\simeq$5-7~R$_*$, with expansion velocities below 10~km~s$^{-1}$. We interpret the pumping mechanism as due to overlaps between $v$=1-0 ro-vibrational lines of SiS and mid-IR lines of C$_2$H$_2$, HCN and their $^{13}$C isotopologues. The large number of overlaps found suggests the existence of strong masers for high-$J$ $v$=0 and $v$=1 SiS transitions, located in the submillimeter range. In addition, it could be possible to find several rotational lines of the SiS isotopologues displaying maser emission. ", "introduction": "The detection of strong maser emission at the frequencies of pure rotational transitions of some molecules is a common phenomenon in circumstellar envelopes (CSE's) of evolved stars \\citep{elitzur_1992,gray_1999}. The maser is usually produced in a small region of the envelope and sometimes provides valuable information on the physical conditions of the emitting region. Due to the different chemistry, masers are produced by different molecules in O- and C-rich stars. In O-rich stars, SiO exhibits strong maser emission in different rotational transitions within several vibrational states, from $v$=1 to 4 \\citep{cernicharo_1993,pardo_1998}. These masers are formed in a region of the CSE very close to the stellar surface and seem to be driven by NIR radiation \\citep{pardo_2004}. In C-rich stars, although SiO is present with similar abundances than in O-rich stars \\citep{schoier_2006}, no SiO maser has been detected. The explanation could be that SiO is formed at $\\simeq$3-5~R$_*$, where the angular dilution of the star is high and the density and temperature lower than in the regions where SiO masers are produced in O-rich stars \\citep{agundez_2006}. In C-rich stars only HCN shows strong maser emission in several pure rotational lines within vibrational states from $\\nu_2$=1 to 4 \\citep{lucas_1989,schilke_2003}. These masers must be formed in the innermost regions of the CSE. SiS has been previously found to show weak maser emission in the $J$=1-0 $v$=0 transition in IRC+10216 \\citep{henkel_1983}. In this letter, we report on the first detection of maser emission from the $J$=11-10, 14-13 and 15-14 transitions in the $v$=0 vibrational state of SiS (hereafter M$_1$, M$_2$, and M$_3$) observed toward the C-rich star IRC+10216. We have also obtained observations of $v$=1 rotational lines which exhibit thermal emission. We propose that overlaps of $v$=1-0 ro-vibrational transitions of SiS with mid-IR lines of C$_2$H$_2$ and HCN could provide the pumping mechanism for these masers as well as higher-$J$ $v$=0 SiS masers in the submillimeter range. This discovery is interesting because this species could play in C-rich stars a role similar to that of SiO in O-rich stars: the energy level pattern of both molecules is similar and it is also formed close to the star, as chemical equilibrium and interferometric observations imply \\citep{bieging_1989}. ", "conclusions": "The $v$=1 rotational lines show single cusped profiles and their relative intensities indicate that the rotational levels are thermally populated. The linewidths correspond to velocities of 9-11~km~s$^{-1}$, lower than the terminal velocity $\\simeq$14.5~km~s$^{-1}$ \\citep{cernicharo_2000}, thus the emission arises from the innermost region of the CSE, between the photosphere and the inner dust formation zone, placed at $\\simeq$5~R$_*$ \\citep{keady_1993}. We can derive the SiS abundance in that region from the $v$=1 lines assuming an uniform sphere at a distance of 180 pc with a radius of 10 R$_*$, illuminated by the central star (R$_*$=5$\\times$10$^{13}$ cm, T$_*$=2300~K), with T$_\\textnormal{\\scriptsize{k}}$=1000~K and n$_{\\textnormal{\\scriptsize{H}}_2}$=1.6$\\times$10$^{9}$~cm$^{-3}$ (mean values for this region derived by \\citealt{fonfria_2006}) and an expansion velocity of 11~km~s$^{-1}$. We solve the statistical equilibrium for SiS considering 100 rotational levels and 3 vibrational states and apply the LVG radiative transfer formalism using the code developed by J. Cernicharo. With the assumed n$_{\\textnormal{\\scriptsize{H}}_2}$, the column density for H$_2$ is $\\simeq$7$\\times$10$^{23}$~cm$^{-2}$ and the derived one for SiS is $\\simeq$5.0$\\times$10$^{18}$~cm$^{-2}$. Hence, the SiS abundance is $\\simeq$7$\\times$10$^{-6}$. This result is compatible with a higher SiS abundance in the innermost CSE (from LTE chemistry models, 3$\\times$10$^{-5}$, \\citealt{agundez_2006}) and an abundance further away of 6.5$\\times$$10^{-7}$ according to observations of $v$=0 rotational lines over the outer CSE by \\citet{bieging_1989}. Most $v$=0 rotational lines show a rounded or slightly double peaked profile with the blue part absorbed by cold SiS through the envelope. However, M$_1$, M$_2$ and M$_3$ show extra emission in the form of narrow peaks (FWHM=1-3~km s$^{-1}$). The velocities of these features, within the $-10$ to 10~km s$^{-1}$ range, indicate that SiS maser emission arises from different regions located between the star and the inner dust formation zone ($r$$\\simeq$5~R$_*$). The lines M$_2$ and M$_3$ have been previously observed by \\citet{sahai_1984} but no maser emission was noticed. This could be due either to the limited sensitivity of their observations or to a time variability of the SiS maser phenomenon. The bottom-center panel of Fig.~\\ref{fig:figure} shows in detail the line profiles of the observed masers. Up to ten maser features labelled ($a$,\\ldots,$j$) are identified. The most complex line profile is that of M$_3$. It is formed by 5 main features: $a$, $d$, $f$, $i$, $j$ with $v$=$-8.8$, $-3.7$, $0.88$, $6.3$ and $8.4$~km s$^{-1}$, respectively. The line profiles show that the strongest peaks are at negative velocities, having their red counterparts rather weak. This behavior was previously found by \\citet{henkel_1983} for the $v$=0 $J$=1-0 line, which mostly consists of a narrow peak (FWHM=0.3~km~s$^{-1}$) centered at $-13.5$~km~s$^{-1}$. The strong asymmetry of the lineshapes can be either due to blanking by the star of the redshifted maser or by amplification of the blueshifted emission by the foreground stellar environment. The strongest observed maser, M$_3$, with T$_\\textnormal{\\scriptsize{MB}}$$\\simeq$60~K (F$_\\textnormal{\\scriptsize{obs}}$$\\simeq$300~Jy; with thermal and non-thermal emission), is weak compared to some SiO masers detected in O-rich stars \\citep*[e.g.][]{cernicharo_1993} or to HCN masers observed in IRC+10216 \\citep{lucas_1989,schilke_2003}. However, M$_1$, M$_2$, and M$_3$ are stronger than the SiS $J$=1-0 maser observed towards IRC+10216 by \\citet{henkel_1983}. The similarity between maser features in M$_2$ and M$_3$ indicates that they may arise from the same regions and produced by the same pumping mechanism; the maser in M$_1$ is probably formed in other regions. Hence, we suggest two possible geometries of the innermost CSE to explain the observed features: ($i$) An onion-like innermost region, where each maser is produced in a shell. This hypothesis is supported by the symmetry of features $a$--$j$ and $d$--$i$. The peaks at extreme velocities, $a$ and $j$, would be produced just in front of and behind the star near the inner dust formation shell ($r$$\\simeq$5~R$_*$) with expansion velocities of $\\simeq$5-11~km~s$^{-1}$. The features $d$ and $i$ would be produced in a similar way but in an inner shell with a lower expansion velocity. Finally, the central peak, $f$, would be formed in a shell very close to the star, with the whole shell contributing to the maser emission. M$_1$ would be produced in a cap-shaped region in front of the star. ($ii$) All the masers are formed in different positions of a clumpy shell. The different features in M$_2$ and M$_3$ would be produced in different regions of the shell: peaks $a$ and $j$ in front of and behind the star and the other peaks ($d$, $f$, and $i$) in different clumps, as occur with the only feature of M$_1$. The classic pumping mechanism for the SiO $v$$>$0 masers observed in O-rich stars resides in the increase of the trapping lifetime (A/$\\tau$)$_{v \\rightarrow v-1}$ with $J$ for $v$$\\rightarrow$$v$-1 transitions, when they become optically thick \\citep{kwan_1974}. Such mechanism produces masers in adjacent rotational lines of the $v$ state, and explains the $v$=1 and 2 SiO masers \\citep{bujarrabal_1981,lockett_1992}. However, the masers observed in rotational transitions of $^{29}$SiO, $^{30}$SiO, and in $v$=3 and 4 of SiO do not show the latter behavior and have been interpreted as due to IR overlaps between ro-vibrational lines of SiO isotopologues \\citep{cernicharo_1991,gonzalezalfonso_1997}. For SiS, the absence of maser emission in $v$=1 rotational lines and the odd $v$=0 pattern also exclude the Kwan \\& Scoville pumping mechanism. This suggests that overlaps of $v$=1-0 ro-vibrational transitions of SiS with those of mode $\\nu_5$ of C$_2$H$_2$ and mode $\\nu_2$ of HCN, could provide the pumping mechanism. C$_2$H$_2$ and HCN are abundant in the inner CSE of IRC+10216 and dominate the 11-14 $\\mu$m spectrum \\citep{fonfria_2006}. Overlaps with these two species have been already proposed by \\citet{sahai_1984} to explain the different profiles of adjacent $J$ lines of SiS. However, the SiS frequencies used in that work were not as accurate ($\\sigma$$\\sim$10$^{-1}$~cm$^{-1}$) as those available today. We have calculated those frequencies from the Dunham coefficients determined by \\citet{sanz_2003}, for which the error of the band center is $<$$10^{-4}$~cm$^{-1}$ ($\\simeq$0.04~km~s$^{-1}$; the relative accuracy of P and R lines is much better). The frequencies of C$_2$H$_2$, H$^{13}$CCH, HCN, and H$^{13}$CN lines have been taken from the HITRAN Database 2004 \\citep{rothman_2005}, with an accuracy better than $10^{-3}$ cm$^{-1}$ ($\\simeq$0.4~km~s$^{-1}$) for C$_2$H$_2$ and H$^{13}$CCH, and $10^{-4}$~cm$^{-1}$ for HCN and H$^{13}$CN. Table~\\ref{tab:frequencies_sis} shows the mid-IR line overlaps of SiS with C$_2$H$_2$, HCN, and their most abundant isotopologues. For the overlap search we selected coincidences within $|\\Delta v|$$<$10~km~s$^{-1}$. However, since the CSE is expanding, every region of the envelope is receding from the others. Hence, if the population of the SiS levels is affected by an overlap with a strong line of other species, the frequency of this overlapping transition must be higher than the SiS one. This would restrict the condition to positive $\\Delta v$. Nevertheless, due to the linewidth, lines at $\\Delta v$$<$0 can overlap the SiS lines. Hence, we have set the negative cutoff to one half of the typical linewidth in the innermost CSE ($\\simeq$5~km~s$^{-1}$; \\citealt{fonfria_2006}). Therefore, our search is restricted to $-2.5\\le\\Delta v\\le 10$~km~s$^{-1}$. All the ro-vibrational SiS lines commented hereafter refer to $v$=1$\\rightarrow$0 transitions and will be labelled with the usual spectroscopic nomenclature R, Q, P (see footnote of Table~\\ref{tab:frequencies_sis}). \\begin{deluxetable}{c@{}c|c@{}c@{}c} \\tabletypesize{\\scriptsize} \\tablewidth{0pc} \\tablecolumns{5} \\tablecaption{Mid-infrared Line Overlaps of SiS with C$_2$H$_2$, HCN, and Their Most Abundant Isotopologues} \\tablehead{\\colhead{Line} & \\colhead{$\\nu$ (cm$^{-1}$)} & \\colhead{Mol.} & \\colhead{Transition} & \\colhead{$\\Delta v$(km/s)}} \\startdata \\multicolumn{5}{c}{\\textbf{Overlaps Involving SiS Levels of Observed Lines}}\\\\[2pt] R 9 & 750.3695 & HCN & $01^{1}0-00^{0}0$ R$_e\\left(12\\right)$ & $-6.6$ \\\\ P10 & 738.2889 & C$_2$H$_2$ & $1^{-1}1^{1}-1^{-1}0^{0}$ R$_f\\left(3\\right)$ & \\phs{}1.2 \\\\ R13 & 752.6431 & C$_2$H$_2$ & $0^{0}1^{1}-0^{0}0^{0}$ R$_e\\left(9\\right)$ & \\phs{}5.9 \\\\[2pt] P14 & 735.7324 & C$_2$H$_2$ & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(38\\right)$ & $-6.1$ \\\\ P15 & 735.0861 & C$_2$H$_2$ & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(36\\right)$ & $-9.9$ \\\\[2pt] \\multicolumn{5}{c}{\\textbf{Overlaps Involving SiS Levels of Unobserved Lines}}\\\\[2pt] P 2 & 743.2624 & C$_2$H$_2$ & $0^{0}1^{1}-0^{0}0^{0}$ R$_e\\left(5\\right)$ & \\phs{}0.6 \\\\ P16 & 734.4368 & C$_2$H$_2$ & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(34\\right)$ & \\phs{}1.1 \\\\ P20 & 731.8114 & C$_2$H$_2$ & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(24\\right)$ & \\phs{}9.5 \\\\ P22 & 730.4815 & C$_2$H$_2$ & $1^{-1}1^{1}-1^{1}0^{0}$ Q$_e\\left(18\\right)$ & $-2.1$ \\\\ R22 & 757.5819 & C$_2$H$_2$ & $1^{1}1^{1}-1^{1}0^{0}$ R$_e\\left(10\\right)$ & \\phs{}0.1 \\\\ R22 & 757.5819 & H$^{13}$CN & $01^{1}0-00^{0}0$ R$_e\\left(17\\right)$ & \\phs{}5.2 \\\\ P23 & 729.8123 & H$^{13}$CCH & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(19\\right)$ & \\phs{}8.7 \\\\ P24 & 729.1402 & C$_2$H$_2$ & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(1\\right)$ & \\phs{}9.4 \\\\ P25 & 728.4654 & H$^{13}$CCH & $0^{0}1^{-1}-0^{0}0^{0}$ Q$_e\\left(7\\right)$ & $-0.0$ \\\\ R25 & 759.1733 & HCN & $01^{1}0-00^{0}0$ R$_e\\left(15\\right)$ & \\phs{}3.6 \\\\ R26 & 759.6977 & C$_2$H$_2$ & $0^{0}1^{1}-0^{0}0^{0}$ R$_e\\left(12\\right)$ & $-1.0$ \\\\ R33 & 763.2816 & H$^{13}$CN & $01^{1}0-00^{0}0$ R$_e\\left(19\\right)$ & \\phs{}6.1 \\\\ P38 & 719.4363 & C$_2$H$_2$ & $1^{1}1^{1}-1^{1}0^{0}$ P$_e\\left(5\\right)$ & \\phs{}4.5 \\\\ R40 & 766.7130 & C$_2$H$_2$ & $0^{0}1^{1}-0^{0}0^{0}$ R$_e\\left(15\\right)$ & \\phs{}4.3 \\\\ R42 & 767.6652 & C$_2$H$_2$ & $1^{1}1^{-1}-1^{1}0^{0}$ R$_e\\left(20\\right)$ & \\phs{}8.7 \\\\ R45 & 769.0697 & C$_2$H$_2$ & $0^{0}1^{1}-0^{0}0^{0}$ R$_e\\left(16\\right)$ & $-1.8$ \\\\ P48 & 712.1720 & C$_2$H$_2$ & $1^{-1}1^{-1}-1^{-1}0^{0}$ P$_f\\left(8\\right)$ & \\phs{}4.7 \\enddata \\tablecomments{Overlaps of SiS with C$_2$H$_2$, H$^{13}$CCH, HCN and H$^{13}$CN found in the mid-IR range $\\left[690,780\\right]$~cm$^{-1}$ with $-2.5\\le\\Delta v\\le 10$ km~s$^{-1}$, where $\\Delta v/c=[\\nu (\\textnormal{X})-\\nu (\\textnormal{SiS})]/ \\nu (\\textnormal{SiS})$ and $J\\le 50$, (corresponding SiS $v$=0 rotational frequencies below 900~GHz). The lines and frequencies at the left correspond to the vibrational transitions $v$=1$\\rightarrow$0 of SiS. The errors on the velocities are $< 0.5$~km~s$^{-1}$. The notation for the C$_2$H$_2$ and H$^{13}$CCH vibrational states involved in the ro-vibrational transitions is $\\nu_4^{\\ell_4}\\nu_5^{\\ell_5}$, whereas for HCN and H$^{13}$CN is $\\nu_1\\nu_2^\\ell\\nu_3$. The parity of the lower level is even ($e$) and odd ($f$). The transitions are labelled as R, P, and Q for $J_{up}-J_{low}=+1$, $-1$, 0.} \\label{tab:frequencies_sis} \\end{deluxetable} In order to qualitatively interpret the effects of IR overlaps on maser emission, we have used the same LVG radiative transfer code modified to account for overlaps, changing the intensity at the overlapping frequency and the escape probability for photons from the overlapped lines. Thus, for some SiS lines, the excitation temperature becomes negative (maser activation) and the brightness temperature is considerably enhanced. M$_2$ and M$_3$ are naturally explained by the overlap of the R$\\left(13\\right)$ line of SiS with the strong C$_2$H$_2$ line $0^01^1-0^00^0$R$_e\\left(9\\right)$ (Table~\\ref{tab:frequencies_sis}). SiS molecules are easily excited from $v$=0 $J$=13 to $v$=1 $J$=14 through R$\\left(13\\right)$ and decay to the $v$=0 $J$=15 via P$\\left(15\\right)$, creating a population inversion between the $v$=0 $J$=13 and 15, and producing maser emission in M$_2$ and M$_3$. M$_1$ may be produced by the overlap of the P$\\left(10\\right)$ SiS line with the strong C$_2$H$_2$ line $1^{-1}1^1-1^{-1}0^0$R$_f\\left(3\\right)$. This overlap can pump SiS molecules from $v$=0 $J$=10 to $v$=1 $J$=9 through P$\\left(10\\right)$, depopulates the $v$=0 $J$=10 level and produces the inversion between $v$=0 $J$=10 and 11. We have also looked for overlaps of $v$=1-0 higher-$J$ SiS lines with C$_2$H$_2$ and HCN transitions to try to predict SiS masers at submillimeter wavelengths. Some of them are shown in the second block of Table~\\ref{tab:frequencies_sis}. They suggest, for example, that a maser could be found in rotational transitions involving the $v$=0 $J$=23 level (likely $J$=24-23 and maybe 23-22), or the $v$=0 $J$=25 and 26 states (perhaps in the $v$=0 $J$=27-26 and maybe 26-25). These overlaps could also produce masers in $v$=1 rotational transitions. There are many overlaps of other SiS isotopologues with lines of C$_2$H$_2$, HCN and $^{28}$Si$^{32}$S. With the adopted criteria for the overlap search (see footnote of Table~\\ref{tab:frequencies_sis}), we have found 91, 93, 76, and 94 coincidences for $^{29}$SiS, $^{30}$SiS, Si$^{34}$S, and Si$^{33}$S, respectively. Consequently, although these species are less abundant than $^{28}$Si$^{32}$S, the population of some levels could be inverted producing maser emission. This study represents the discovery of three new SiS masers and should be complemented with future observations of higher-$J$ $v$=0 and 1 rotational transitions. Furthermore, a detailed multi-molecule non-local radiative transfer model would help to understand the dust formation region and the role of SiS in C-rich evolved stars." }, "0710/0710.0600_arXiv.txt": { "abstract": "% We describe the status of a project whose main goal is to detect variability along the extreme horizontal branch of the globular cluster NGC~6752. Based on Magellan 6.5m data, preliminary light curves are presented for some candidate variables. By combining our time-series data, we also produce a deep CMD of unprecedented quality for the cluster which reveals a remarkable lack of main sequence binaries, possibly pointing to a low primordial binary fraction. ", "introduction": "Among field B-type subdwarf (sdB) stars, three types of non-radial pulsators have so far been detected, namely: \\begin{itemize} \\item EC 14026 (sdBV) stars: these are $p$-mode pulsators whose temperatures fall in the range between 29,000 and 36,000~K. Their periods are typically found in the range 100-200~sec, and their amplitudes cover the range from 0.4 to 25\\%. \\item PG1716+426 (``Betsy'') stars: these are $g$-mode pulsators, with temperatures in the range between 25,000 and 30,000~K. Their periods are much longer, typically falling in the range between 2000 and 9000~sec, and their amplitudes are smaller than 0.5\\%. \\item Hybrid stars: these are stars that present simultaneous $p$- and $g$- mode oscillations. At present, only two examples have been reported in the literature \\citep{abea05,roea05,ssea06}. \\end{itemize} Performing asteroseismology on these stars holds the promise to unveil their innermost secrets, thus providing an exciting new route toward the solution of the so-called ``second-parameter problem'' \\citep[][and references therein]{mc05}. The great promise of the technique notwithstanding, such variables have never been detected in previous searches in globular clusters \\citep[e.g.,][]{mrea06}. Accordingly, the main purpose of the present study is to perform a new search for this type of variables in the relatively nearby southern globular cluster NGC~6752, which contains a very long blue HB ``tail,'' and thus many potential candidates for the three aforementioned variability types. ", "conclusions": "Our analysis of time-series observations collected with the Magellan 6.5m MagIC camera already reveals a few intriguing variable candidates, including at least one on the hot extension of the HB, in two small fields located away from the cluster center. We have recently acquired extensive time-series observations using IMACS over a much larger field. These data will help confirm the nature of the suspected variables, and will presumably reveal a host of other faint/low-amplitude variables in NGC~6752. In addition, the new data will allow us to map the MS binary fraction in the cluster as a function of radius, which will reveal, for the first time, how in detail the binary fraction decreases as a function of radius in a globular star cluster." }, "0710/0710.2433_arXiv.txt": { "abstract": "We investigate the wave effect in the gravitational lensing by a black hole with very tiny mass less than $10^{-19} M_{\\rm sun}$ (solar mass), which is called attolensing, motivated by a recent report that the lensing signature might be a possible probe of a modified gravity theory in the braneworld scenario. We focus on the finite source size effect and the effect of the relative motion of the source to the lens, which are influential to the wave effect in the attolensing. Astrophysical condition that the lensed interference signature can be a probe of the modified gravity theory is demonstrated. The interference signature in the microlensing system is also discussed. ", "introduction": "Many physicists have drawn attention to extra dimensional physics for several years due to recent development in testing Randall-Sundrum type II (RS-II) scenario \\cite{RSII}. In this scenario they considered a four dimensional positive-tension brane embedded in five dimensional AdS bulk which allows us to reconsider our understanding about the history of our universe in the early stages. Some investigations have been carried out to modify the existence of primordial black holes (PBH)s in this RS-II scenario that the life time of five dimensional PBHs against the Hawking radiation becomes longer compared with the standard PBH in four dimensions \\cite{GCL}. This is because the five dimensional feature becomes significant in the black hole with very tiny mass. The ratio of the life time against the Hawking radiation of such five dimensional PBH to that of the four dimensional PBH may be estimated, $lM_4/l_4M$, where $l$ is the AdS radius of the braneworld model, $l_4$ and $M_4$ are the four dimensional Planck length and mass, respectively, and $M$ is the black hole mass \\cite{GCL}. Since the braneworld PBHs can live longer, then it is possible that such black holes still exist and spread out in our universe. Therefore, it is natural to consider the possibility of the gravitational lensing phenomenon by such the black hole. More recently, the authors \\cite{KPIII} investigated the wave effect in gravitational lensing by the black hole with the very tiny mass smaller than $10^{-19} M_{\\rm sun}$, where $M_{\\rm sun}$ is the solar mass, which is called attolensing. They showed that the interference signature in the energy spectrum in gamma ray burst due to the attolensing might be a possible probe of the modified gravity theory. In the standard general relativity, PBH with the mass smaller $10^{-18} M_{\\rm sun}$ will be evaporated through the Hawking radiation within the cosmic age. Then, the detection of such the interference signature would be a probe of extra dimension of our universe. In the present paper, we consider the two effects in the lensing phenomenon that are influential for measurement of the interference signature in the energy spectrum in the attolensing: One is the finite source size effect. The other is the effect of the relative motion of the black hole (lens object) to the source. We demonstrate the condition that these effects become influential. This is one of the astrophysical conditions that the attolensing can be a probe of the modified gravity theory. Throughout this paper, $ H_0 = 72$ km s$^{-1}$ Mpc$^{-1}$ is the Hubble parameter, and we use the unit in which the light velocity equals 1. ", "conclusions": "We investigated the condition that the finite source size effect and the relative motion of the source becomes substantial in the wave effect of the gravitational lensing. The condition is expressed by the formula (\\ref{condtwo}), whose physical meaning is that the difference of the phase between the two light paths from $y_{\\rm max}$ and $y_{\\rm min}$ is larger than $2\\pi$. We have shown that the finite source size effect is important in the attolensing by the black hole at the cosmological distance. Also the relative motion of the source to the lens can be influential. For the attolensing by the black hole at the galactic distance, the constraint is relaxed by the factor $\\sqrt{D_S/D_L}\\sim 10^{3}$. However, we should also note that the detection of the attolensing is limited in practice \\cite{KPIII}, even when the finite source size is not taken into account. In general, the signature of the interference becomes remarkable when $w\\sim 1$, i.e., \\begin{eqnarray} w \\sim 0.3\\times (1+z_L)\\left({h\\nu\\over 100{\\rm MeV}}\\right) \\left({M\\over 10^{-19}M_{\\rm sun}}\\right)\\sim 1. \\label{condfirst} \\end{eqnarray} In the case $w\\simlt 1$, the amplification due to the lensing becomes negligible, because the wavelength of the light becomes larger than the size of deflector. Combining this and the condition (\\ref{condtwo}), we may conclude that the condition for the possible observation of the interference signature in the gravitational lensing needs \\begin{eqnarray} 1\\simlt w \\simlt {\\pi \\over |y_{\\rm min}-y_{\\rm max}|}. \\label{finalcond} \\end{eqnarray} Therefore, this means $|y_{\\rm min}-y_{\\rm max}|\\simlt \\pi$, the angular size of the source must be less than the Einstein radius, is always necessary for a possible observation of the interference signature. It will be useful to demonstrate the region satisfying (\\ref{finalcond}) clearly. In Figure 8, the dashed line is $w=1$, and the solid line is $w|y_{\\rm min}-y_{\\rm max}| =\\pi$, where we fixed $\\hat R=10^3$ km and $D_{LS}D_S/D_L=H_0^{-1}$. The shaded region satisfies the condition (\\ref{finalcond}). It might also be interesting to consider whether other physical system satisfy the condition (\\ref{finalcond}) or not. Figure 9 shows a region satisfying the condition with the parameter associated with the microlensing. Similar to Figure 8, the shaded region satisfies the condition. Here the dashed line in Figure 9 is $w=1$, and the solid line is $w|y_{\\rm min}-y_{\\rm max}| =\\pi$, where we fixed $\\hat R=7\\times 10^5$ km (solar radius) and $D_{LS}D_S/D_L=30$ kpc. The point of the intersection of these two lines is \\begin{eqnarray} && \\left({\\nu\\over {\\rm GHz}}\\right)=8.9\\times 10^2 \\left({\\hat R\\over 7\\times 10^5{\\rm km}} \\right)^{-2} \\left({30{\\rm kpc}\\over D_{LS}D_S/D_L}\\right)^{-1} \\\\ && \\left({M\\over M_{\\rm sun}}\\right)=0.9\\times 10^{-8} \\left({\\hat R\\over 7\\times 10^5{\\rm km}} \\right)^{2} \\left({30{\\rm kpc}\\over D_{LS}D_S/D_L}\\right). \\end{eqnarray} Thus the microlensing can be a possible system that satisfies the condition (\\ref{finalcond}). Especially, for the microlensing by an Earth like planet $M\\sim 10^{-5} M_{\\rm sun}$, the relevant range of the frequency is $1$ GHz $\\sim 100$ GHz. Then, the measurement of the microlensing event through the frequency might be relevant to the interference signature. However, it will be very difficult to detect the signal because the stars at $1\\sim 100$ GHz frequency band at the galactic distance is very dark in general. \\vspace{0.3cm}" }, "0710/0710.5355_arXiv.txt": { "abstract": "{ The ANTARES Collaboration is building a high-energy neutrino telescope at 2500 m depth in the Mediterranean Sea. The experiment aims to search for high-energy cosmic neutrinos through the detection of Cerenkov light induced by muons and showers resulting from neutrino interactions with the surrounding medium. The detector will consist of a three-dimensional array of 900 optical modules housing photomultipliers. It will be composed of 12 strings, 5 of them being already in operation since January 2007. The muon track is reconstructed from the arrival time and the charge of the signals obtained from the photomultipliers, whose positions are known by means of an acoustic positioning system. The reconstruction strategies include several steps among which there are: optical background filtering, algorithms for first estimations of the track parameters, and a final fit aiming to reach an angular resolution better than 0.3 degree above 10 TeV in the full detector. Different reconstruction strategies will be presented and their application to the present real data analysis will be reviewed.} \\begin{document} ", "introduction": "When an ultra-relativistic particle ($\\beta \\simeq 1$) moves in a medium, Cerenkov light is emitted at an angle depending on the refraction index $n$ of the medium. In case of sea water $n$ is about $1.34$, and thus the Cerenkov emission angle becomes \\begin{equation} cos (\\theta_C) = \\frac{1}{n} \\simeq cos(42^\\circ). \\end{equation} The emitted Cerenkov photons are detected by an array of photomultipliers installed at the bottom of the sea. Since January 2007 the ANTARES detector is a full three-dimensional array of photomultipliers consisting of 5 strings detecting muons at a depth of 2475 meters below the sea level. ANTARES 10-inch photomultipliers are housed in pressure resistant glass spheres called optical modules (OM). The 5-line detector data taking has allowed the Collaboration to tune the various processes needed to collect data, and muon events have been detected on top of the optical background (60-100 kHz most of the time). The calibration system has been proven to be efficient, and the knowledge of the charge, of the arrival time of the signals and the positioning of the OMs has allowed the first reconstruction codes to be tested in realistic conditions. The status of the experiment is discussed in \\cite{antoine}. Preliminary data studies show that a flux of atmospheric downward-going muons is triggering the detector at a rate of about 1 Hz and that upgoing atmospheric neutrino candidates have been identified. Among the various muon reconstruction algorithms under study, two different alternative methods are presented in this paper. Other well developed reconstruction methods are described in \\cite{aart}, \\cite{carmona}, and the implementation of the {\\sl{simulated annealing}} algorithm \\cite{simann} is under study. A discussion on event reconstruction techniques for Cerenkov neutrino telescopes can also be found in \\cite{amanda}. ", "conclusions": "" }, "0710/0710.5163_arXiv.txt": { "abstract": "We report the discovery of 11 new cataclysmic variable (CV) candidates by the Isaac Newton Telescope (INT) Photometric H$\\alpha$ Survey of the northern Galactic plane (IPHAS). Three of the systems have been the subject of further follow-up observations. For the CV candidates IPHAS J013031.90+622132.4 and IPHAS J051814.34+294113.2, time-resolved optical spectroscopy has been obtained and radial-velocity measurements of the H$\\alpha$ emission-line have been used to estimate their orbital periods. A third CV candidate (IPHAS J062746.41+ 014811.3) was observed photometrically and found to be eclipsing. All three systems have orbital periods above the CV period-gap of 2--3\\,h. We also highlight one other system, IPHAS J025827.88+635234.9, whose spectrum distinguishes it as a likely high luminosity object with unusual C and N abundances. ", "introduction": "\\label{intro} Cataclysmic variables (CVs) are semi-detached interacting binary systems containing a white dwarf (WD) primary and a late-type main-sequence secondary. The secondary fills its Roche lobe, and matter is transferred to the primary through the inner Lagrangian point. The mass-transfer process in CVs means that line emission is a common observational feature of the majority of CVs, especially in the Balmer series. Observed lines may originate from optically thin or irradiated parts of the accretion disc (if present, \\citealt{1980ApJ...235..939W}). Line emission may also be produced in the accretion stream and on the irradiated face of the secondary star. \\citet{1995cvs..book.....W} provides a comprehensive review of CVs. Population synthesis models suggest that the intrinsically faintest low accretion rate systems should dominate the Galactic CV population and should be found predominantly at short orbital periods, that is, below the well-known CV ``period gap'' between 2 hrs and 3 hrs \\citep{1993A&A...271..149K, 1997MNRAS.287..929H}. This dominant population of faint, short-period CVs has proven quite elusive. This is partly because most known CVs have been discovered with techniques that are biased against detecting CVs with low mass accretion rates. For example, all flux-limited surveys with relatively bright limiting magnitudes and also variability searches (e.g. for dwarf nova outbursts) are intrinsically biased against the detection of faint, short-period CVs with potentially long inter-outburst recurrence times (such as WZ Sge). For a closer look at the period distributions of CVs found with different discovery methods, see \\citet{2005ASPC..330....3G}. Recently \\citet{2007MNRAS.374.1495P} have shown that the relative dearth of short-period CVs is not just due to selection effects, implying a serious flaw in our understanding of CV evolution (see also \\citealt{2006A&A...455..659A}). Nevertheless, selection effects must be at least partly responsible for the lack of known faint short-period CVs, and so many such systems still remain to be found. Given the ubiquity of line emission amongst CVs, searches for objects displaying H$\\alpha$ emission offer a powerful way to find new CVs. Examples of such CV searches in the Galactic field include those of \\citet{2002ASPC..261..190G}, \\citet{2005A&A...443..995A}, \\citet{2002A&A...395L..47H}, and the Chandra Multiwavelength Plane (ChaMPlane) survey (\\citealt{2005ApJ...635..920G}, \\citealt{2005ApJS..161..429Z} and \\citealt{2006ApJS..163..160R}). Similar searches have also been performed in globular clusters have been performed by \\citet{1995ApJ...439..695C}, \\citet{1995ApJ...455L..47G}, \\citet{1996ApJ...473L..31B} and \\citet{2000ApJ...532..461C}. The Isaac Newton Telescope (INT) Photometric H$\\alpha$ Survey of the northern Galactic plane (IPHAS) is currently surveying the Milky Way in broad-band Sloan $r^\\prime$ and $i^\\prime$ and narrow-band H$\\alpha$ and provides an excellent data base for a detailed CV search at low Galactic latitudes. The survey goes to a depth of $r^\\prime\\simeq20$\\,mag and covers the latitude range $-5^o < b < +5^o$. A detailed introduction to the survey, including the transmission profiles of the filters it uses, is given by \\citet{2005MNRAS.362..753D}. What makes IPHAS particularly promising for finding CVs is that, empirically, the intrinsically faintest, low mass transfer rate ($\\dot{M}$) systems tend to have the largest Balmer line equivalent widths (EWs; \\citealt{1984ApJS...54..443P}; \\citealt{2006MNRAS.369..581W}). Unless the inverse relationship between Balmer EW and orbital period breaks down for the faintest, short-period systems, emission-line surveys should be very good at finding low-$\\dot{M}$ CVs. IPHAS should then be an excellent way of detecting a large population of faint, short-period CVs. In this paper we announce the discovery of eleven new CV candidates from the IPHAS survey data. We present spectroscopic follow-up observations of two of these CVs candidates and determine their orbital periods. We also present the results of time-series photometry of another CV candidate which establishes it as a long period, eclipsing, system. ", "conclusions": "\\label{conc} Spectroscopy of several objects from the IPHAS H$\\alpha$ excess catalogue has led to the discovery of 11 new CV candidates by virtue of their strong H$\\alpha$ emission. The identification spectra of the CVs candidates include cases where the secondary star is particularly prominent (IPHAS J1856), and examples of double-peaked emission lines (for example IPHAS J0518). One particular candidate (IPHAS J0258) has similarities to the extreme objects V Sge and QU Car. Time-resolved optical spectroscopy of two of the new CVs (IPHAS J0130 and IPHAS J0518) has been obtained using CAFOS on the 2.2\\,m telescope at the Calar Alto Observatory. Radial-velocity measurements of the H$\\alpha$ emission-line were used to determine their orbital periods. Periodograms and a Monte Carlo analysis were used to estimate the orbital period of $P_{orb}=0.130061 \\pm 0.000001$\\,d for IPHAS J0130. The periodogram obtained from the radial-velocity data of IPHAS J0518 reveals four peaks which could be the orbital period of the system: 0.2383, 0.2203, 0.2595 and 0.2049\\,d, each with an uncertainty of 0.0007\\,d. Time-series photometry of IPHAS J0627 using the SAAO 1.9\\,m telescope found the system to be a long-period eclipser with four possible orbital periods: $1.020 \\pm 0.002$\\,d, $0.5101 \\pm 0.0008$\\,d, $0.3401 \\pm 0.0006$\\,d and $0.2551 \\pm 0.0004$\\,d. Further observations are necessary to obtain accurate values of the orbital period of all three systems, and to find accurate binary parameters." }, "0710/0710.4954_arXiv.txt": { "abstract": "We study the signal for the detection of quasi-stable supersymmetric particle produced in interactions of cosmogenic neutrinos. We consider energy loss of high energy staus due to photonuclear and weak interactions. We show that there are optimal nadir angles for which the stau signal is a factor of several hundred larger than muons. We discuss how one could potentially eliminate muon background by considering the energy loss of muons in the detector. We also show results for the showers produced by weak interactions of staus that reach the detector. ", "introduction": "Ultrahigh energy cosmic neutrinos could potentially probe physics beyond the Standard Model \\cite{ringwald}. Interactions of UHE neutrinos ($E_\\nu \\geq 10^{17}$eV) with nucleons probe center of mass energies above 14 TeV. Some fraction of these neutrinos may produce supersymmetric particles or some other exotic particles. These processes are suppressed relative to standard model processes, however, in some models interesting signals may arise from supersymmetric particles with long lifetimes. In most SUSY scenarios, particles produced in high energy collisions decay immediately into the lightest one and are thus hard to detect. However in some low scale supersymmetric models in which gravitino is the lightest supersymmetric particle (LSP) and R-parity is conserved, the next-to-lightest particle (NLSP) is the charged superpartner of the right-handed tau, the stau \\cite{susy_models}. Due to its weak coupling to the gravitino, the stau is a long-lived particle in these models. For the supersymmetric breaking scale, $\\sqrt F >5\\times10^6$ GeV, the long-lived stau could travel distances of the order of $10^4$ km before decaying into the gravitino. The distance that staus travel before decaying depends on the gravitino mass (or equivalently on the supersymmetry breaking scale) and the stau mass. Limits on the stau mass of about 100 GeV come from its non-observation in accelerator experiments \\cite{stmass1,stmass2,stmass3,stmass4}. Recently it was proposed that staus produced in high energy neutrino interactions, where neutrinos originate in astrophysical sources, might be detectable in neutrino telescopes \\cite{ABC,Ahlers}. The cross section for the production of staus in neutrino-nucleon scattering \\cite{ABC} is several orders of magnitude smaller than the neutrino charged-current or neutral-current cross section \\cite{gqrs}. However, once produced, the long-lived staus have the potential to travel through the earth without decaying and thus open up a possibility to be detected in neutrino telescopes. The long range of staus could potentially compensate for the suppression in the production cross section by increasing the effective detector volume and therefore enhancing the signal. The detection of staus depends on the stau lifetime and range, so it is important to determine the energy loss as it traverses the earth. The details of the range depend in part on the supersymmetry breaking and how the quasi-stable stau particle is comprised of the SUSY partners of the right-handed and left-handed taus. The electromagnetic energy loss has been shown to have the largest contribution from photonuclear interactions for stau energies between $10^6$-$10^{12}$ GeV, resulting in a range of $10^4$ km.w.e. for masses of the order of a few hundred GeV \\cite{RSS}. Weak interactions may come into play as well. The stau range has been shown to be sensitive to the mixing angle of right-handed and left-handed staus. When the mixing is maximal, weak interactions act to suppress the range at energies above $\\sim10^9$ GeV \\cite{RSU1}, however, their weak interactions have the potential to produce signals in neutrino detectors such as the Antarctic Impulse Transient Array (ANITA) \\cite{anita} and the Antarctic Ross Iceshelf Antenna Neutrino Array (ARIANNA) \\cite{arianna}. The high energies required for stau production lead us to focus on the production of staus in interactions of cosmogenic neutrinos as they traverse the Earth and/or in the detector. These neutrinos originate from cosmic ray protons interacting with the cosmic microwave background, $$ p\\gamma(3{\\rm K}) \\rightarrow \\Delta \\rightarrow N\\pi$$ followed by charged pion, muon and neutron decays. This flux is guaranteed as cosmic ray fluxes are measured as well as the 3K microwave background. We use a conservative cosmogenic neutrino flux evaluated by Engel, Seckel and Stanev (ESS) in Ref. \\cite{ess}. They evaluate the neutrino flux associated with the measured cosmic ray flux by tracing back cosmic ray propagation through the background radiation. Depending on the cosmological evolution assumed, the overall normalization of this flux has an uncertainty of about a factor of four. In addition, neutrinos could be produced at the sources of the high energy cosmic rays and those are not included in the evaluation of ESS neutrino flux. Thus, the ESS neutrino flux is a conservative estimate of the cosmogenic flux. The cosmogenic neutrino flux, when neutrino flavor oscillations are not included, peaks at high energies, around $10^8$ GeV, and thus it is in the energy range where ANITA and ARIANNA have very good sensitivity \\cite{anita,arianna}. The neutrino flavor ratio for cosmogenic neutrinos deviates from the common 2:1 ratio, due to the neutron decay contribution to electron neutrinos \\cite{ess}. This implies that one needs to consider three flavor oscillations when considering the cosomogenic neutrino flux that arrives at the Earth \\cite{us_JMRS}. We consider cosmogenic neutrinos, their propagation, stau production, and subsequent energy loss as it traverses the earth, for a region of parameter space where the staus do not decay over the distances required. We compare the resulting stau flux when there is no mixing between right-handed and left-handed stau and when there is maximal mixing. We consider muon-like signals (charged tracks) produced by staus and its associated background. We discuss the potential for eliminating the background by measuring the energy loss, which requires large volume detectors. Finally we discuss the showers produced in the ice due to stau interactions and its background from neutrino-induced showers. ", "conclusions": "We have studied signals of staus produced in interactions of cosmogenic neutrinos. We have considered two types of signals, muon-like charged tracks and showers. We have focused on low scale supersymmetric models that have stau as NLSP, which decays into the lightest SUSY particle, the gravitino. For a sufficiently large scale of supersymmetry breaking, the stau has a very long lifetime. Our focus has been on cosmogenic neutrino fluxes and their associated stau production in the Earth. Energy losses, both through electromagnetic and weak interactions, are important in evaluating stau signals. The energy loss of staus, however, is relatively small in comparison with muons. Thus, for some nadir angles, the stau flux is much larger than the muon flux produced in neutrino charged-current interactions. The enhancement of the stau flux is larger from an input cosmogenic neutrino flux than for the Waxman-Bahcall neutrino flux \\cite{WB}. This is because the cosmogenic neutrino flux is peaked at energies of about $10^8$ GeV, while the WB flux is characterized by a steep power law with index of two. The large ratio of staus to muons from cosmogenic neutrinos is encouraging for experimental detection, but in order to see this signal one needs to be able to distinguish between staus and muons. Using the average energy loss per unit distance is not a good way to distinguish staus and muons, since the scaling of the energy loss parameter $\\beta$ has the effect of making a high-energy stau look like a lower energy muon. We have proposed a way to distinguish between stau and muon tracks by measuring the energy loss of muons via their interactions in the ice, and to use this method to reduce the background. We also considered showers produced by staus interacting in the ice via charged-current interactions. The backgrounds for this signal are showers induced directly by neutrinos that reach the detector and interact inside the detector via charged-current or neutral-current interactions. The only way that staus would produce showers in the ice is if there is a weak mixing, but this process also contributes to reducing stau range. These effects combine with the small stau production probability to give fluxes of attenuated staus that are several orders of magnitude less than attenuated neutrino fluxes. This small stau to neutrino ratio translates directly to shower rates. In addition to weak interactions, another possibility for shower production would be stau decays in the detector. For the parameter space considered here, with long-lived staus, the signal from decays is suppressed relative to their weak interactions. In conclusion, stau signals at high energies are best identified by muon-like tracks. The most important detection issue is distinguishing between staus and muons, which may be possible by looking at the incremental electromagnetic energy loss as the charged particle moves through the detection volume. With very large volumes, there is a potential for detection of staus with future large neutrino telescopes." }, "0710/0710.3868_arXiv.txt": { "abstract": "Two dimensional hydrodynamical disks are nonlinearly unstable to the formation of vortices. Once formed, these vortices essentially survive forever. What happens in three dimensions? We show with incompressible shearing box simulations that in 3D a vortex in a short box forms and survives just as in 2D. But a vortex in a tall box is unstable and is destroyed. In our simulation, the unstable vortex decays into a transient turbulent-like state that transports angular momentum outward at a nearly constant rate for hundreds of orbital times. The 3D instability that destroys vortices is a generalization of the 2D instability that forms them. We derive the conditions for these nonlinear instabilities to act by calculating the coupling between linear modes, and thereby derive the criterion for a vortex to survive in 3D as it does in 2D: {\\it the azimuthal extent of the vortex must be larger than the scale height of the accretion disk}. When this criterion is violated, the vortex is unstable and decays. Because vortices are longer in azimuthal than in radial extent by a factor that is inversely proportional to their excess vorticity, a vortex with given radial extent will only survive in a 3D disk if it is sufficiently weak. This counterintuitive result explains why previous 3D simulations always yielded decaying vortices: their vortices were too strong. Weak vortices behave two-dimensionally even if their width is much less than their height because they are stabilized by rotation, and behave as Taylor-Proudman columns. We conclude that in protoplanetary disks weak vortices can trap dust and serve as the nurseries of planet formation. Decaying strong vortices might be responsible for the outwards transport of angular momentum that is required to make accretion disks accrete. ", "introduction": "Matter accretes onto a wide variety of objects, such as young stars, black holes, and white dwarfs, through accretion disks. In highly ionized disks magnetic fields are important, and they trigger turbulence via the magnetorotational instability \\citep{BH98}. However, many disks, such as those around young stars or dwarf novae, are nearly neutral \\citep[e.g.,][]{SMUN00,GM98}. In these disks, the fluid motions are well described by hydrodynamics. Numerical simulations of hydrodynamical disks in two-dimensions---in the plane of the disk---often produce long-lived vortices \\citep{GL99,UR04,JG05}. If vortices really exist in accretion disks, they can have important consequences. First and foremost, they might generate turbulence. Since turbulence naturally transports angular momentum outwards\\footnote{ Energy conservation implies that turbulence transports angular momentum outwards; see \\S \\ref{sec:pseudo}. Nonetheless, if an external energy source (e.g., the radiative energy from the central star) drives the turbulence, then angular momentum could in principle be transported inwards. }, as is required for mass to fall inwards, it might be vortices that cause accretion disks to accrete. Second, in disks around young stars, long-lived vortices can trap solid particles and initiate the formation of planets \\citep{BS95}. Why do vortices naturally form in 2D simulations? Hydrodynamical disks are stable to linear perturbations. However, they are nonlinearly unstable, despite some claims to the contrary in the astrophysical literature. In two dimensions, the incompressible hydrodynamical equations of a disk are equivalent to those of a non-rotating linear shear flow \\citep[e.g.,][hereafter L07]{L07}. And it has long been known that such flows are nonlinearly unstable (\\mycite{Gill65}; \\mycite{LK88}; L07). This nonlinear instability is just a special case of the Kelvin-Helmholtz instability. Consider a linear shear flow extending throughout the $x$-$y$ plane with velocity profile $\\bld{v}=-qx\\bld{\\hat{y}}$, where $q>0$ is the constant shear rate, so that $-q$ is the flow's vorticity. (In the equivalent accretion disk, the local angular speed is $\\Omega=2q/3$.) This shear flow is linearly stable to infinitesimal perturbations. But if the shear profile is altered by a small amount, the alteration can itself be unstable to infinitesimal perturbations. To be specific, let the alteration be confined within a band of width $\\Delta x$, and let it have vorticity $\\omega=\\omega(x)$ (with $|\\omega|\\lesssim q$), so that it induces a velocity field in excess of the linear shear with components $u_y\\sim \\omega\\Delta x$ and $u_x=0$. Then this band is unstable to infinitesimal nonaxisymmetric (i.e. non-stream-aligned) perturbations provided roughly that \\be \\left|k_y \\right| \\lesssim {1\\over q}{|\\omega|\\over \\Delta x} \\ \\ \\Rightarrow {\\rm 2D\\ instability} \\label{eq:2dinst} \\ee where $k_y$ is the wavenumber of the nonaxisymmetric perturbation.\\footnote{ More precisely, the necessary and sufficient condition for instability in the limit $|\\omega|\\ll q$ is that $|k_y|<{1\\over 2 q}\\int_{-\\infty}^\\infty {d\\omega/dx\\over x-x_0}dx$, where $x_0$ is any value of $x$ at which $d\\omega/dx=0$ \\citep[][L07]{Gill65,LK88}. For arbitrarily large $\\omega$, Rayleigh's inflection point theorem and Fj\\o rtoft's theorem give necessary (though insufficient) criteria for instability \\citep{DR04}. The former states that for instability, it is required that $d\\omega/dx=0$ somewhere in the flow, i.e. that the velocity field must have an inflection point. \\cite{Love99} generalize Rayleigh's inflection point theorem to compressible and nonhomentropic disks. \\label{foot:inst} } For any value of $|\\omega|$ and $\\Delta x$, the band is always unstable to perturbations with long enough wavelength. Remarkably, instability even occurs when $|\\omega|$ is infinitesimal. Hence we may regard this as a true nonlinear instability. \\cite{BH06} assert that detailed numerical simulations have not shown evidence for nonlinear instability. The reason many simulations fail to see it is that their boxes are not long enough in the $y$-direction to encompass a small enough non-zero $|k_y|$. In two dimensions, the outcome of this instability is a long-lived vortex (e.g., L07). A vortex that has been studied in detail is the Moore-Saffman vortex, which is a localized patch of spatially constant vorticity superimposed on a linear shear flow \\citep{Saffman95}. When $|\\omega|\\lesssim q$, where $\\omega$ here refers to the spatially constant excess vorticity within the patch, and when the vorticity within the patch ($\\omega-q$) is stronger than that of the background shear, then the patch forms a stable vortex that is elongated in $y$ relative to $x$ by the factor \\be {\\Delta y\\over \\Delta x} \\sim {q\\over |\\omega|} \\ . \\label{eq:ms} \\ee This relation applies not only to Moore-Saffman vortices, but also to vortices whose $\\omega$ is not spatially constant. It may be understood as follows. A patch with characteristic excess vorticity $\\sim\\omega$ and with $\\Delta y\\gg \\Delta x$ induces a velocity field in the $x$-direction with amplitude $u_x\\sim|\\omega|\\Delta x$, independent of the value of $\\Delta y$ (e.g., \\S 6 in L07). As long as $|\\omega|\\lesssim q$, the $y$-velocity within the vortex is predominantly due to the background shear, and is $\\sim q\\Delta x$. Therefore the time to cross the width of the vortex is $t_x\\sim \\Delta x/u_x \\sim 1/|\\omega|$, and the time to cross its length is $t_y\\sim \\Delta y/(q\\Delta x)$. Since these times must be comparable in a vortex, equation (\\ref{eq:ms}) follows. Equation (\\ref{eq:ms}) is very similar to equation (\\ref{eq:2dinst}). The 2D instability naturally forms into a 2D vortex. Futhermore, the exponential growth rate of the instability is $\\sim |\\omega|$, which is comparable to the rate at which fluid circulates around the vortex. More generally, an arbitrary axisymmetric profile of $\\omega(x)$ tends to evolve into a distinctive banded structure. Roughly speaking, bands where $\\omega<0$ contain vortices, and these are interspersed with bands where $\\omega>0$, which contain no vortices. (Recall that we take the background vorticity to be negative; otherwise, the converse would hold.) The reason for this is that only regions that have $\\omega<0$ can be unstable, as may be inferred either from the integral criterion for instability given in footnote \\ref{foot:inst}, or from Fj\\o rtoft's theorem. For more detail on vortex dynamics in shear flows, see the review by \\cite{Marcus93}. What happens in three dimensions? To date, numerical simulations of vortices in 3D disks have been reported in two papers. \\cite{BM05b} initialized their simulation with a Moore-Saffman vortex, and solved the anelastic equations in a stratified disk. They found that this vortex decayed. As it decayed, new vortices were formed in the disk's atmosphere, two scale heights above the midplane. The new vortices survived for the duration of the simulation. \\cite{SSG06} performed both 2D and 3D simulations of the compressible hydrodynamical equations in an unstratified disk, initialized with large random fluctuations. They found that whereas the 2D simulations produced long-lived vortices, in three dimensions vortices rapidly decayed. Intuitively, it seems clear that a vortex in a very thin disk will behave as it does in 2D. And from the 3D simulations described above it may be inferred that placing this vortex in a very thick disk will induce its decay. Our main goal in this paper is to understand these two behaviors, and the transition between them. A crude explanation of our final result is that vortices decay when the 2D vortex motion couples resonantly to 3D modes, i.e., to modes that have vertical wavenumber $k_z\\ne 0$. As described above, a vortex with excess vorticity $|\\omega|$ has circulation frequency $\\sim|\\omega|$, and $k_y/k_x\\sim |\\omega|/q$, where $k_x$ and $k_y$ are its ``typical'' wavenumbers. Furthermore, it is well-known that the frequency of axisymmetric ($k_y=0$) inertial waves is $\\Omega k_z/\\sqrt{k_x^2+k_z^2}$ (see eq. [\\ref{eq:axi3d}]). Equating the two frequencies, and taking the $k_x$ of the 3D mode to be comparable to the $k_x$ of the vortex, as well as setting $q=3\\Omega/2$ for a Keplerian disk, we find \\be k_z\\sim k_y \\label{eq:res} \\ee as the condition for resonance. Therefore a vortex with length $\\Delta y$ will survive in a box with height $\\Delta z\\lesssim \\Delta y$, because in such a box all 3D modes have too high a frequency to couple with the vortex, i.e., all nonzero $k_z$ exceed the characteristic $k_y\\sim 1/\\Delta y$. But when $\\Delta z\\gtrsim \\Delta y$, there exist $k_z$ in the box that satisfy the resonance condition (\\ref{eq:res}), leading to the vortex's destruction. This conclusion suggests that vortices live indefinitely in disks with scale height less than their length ($h\\lesssim \\Delta y$) because in such disks all 3D modes have too high a frequency for resonant coupling. This conclusion is also consistent with the simulations of \\cite{BM05b} and \\cite{SSG06}. Both of these works initialized their simulations with strong excess vorticity $|\\omega|\\sim q$, corresponding to nearly circular vortices. Both had vertical domains that were comparable to the vortices' width. Therefore both saw that their vortices decayed. Had they initialized their simulations with smaller $|\\omega|$, and increased the box length $L_y$ to encompass the resulting elongated vortices, both would have found long-lived 3D vortices. \\citeauthor{BM05b}'s discovery of long-lived vortices in the disk's atmosphere is simple to understand because the local scale height is reduced in inverse proportion to the height above the midplane. Therefore higher up in the atmosphere the dynamics becomes more two-dimensional, and a given vortex is better able to survive the higher it is.\\footnote{ However, \\cite{BM05b} also include buoyancy forces in their simulations, which we ignore here. How buoyancy affects the stability of vortices is a topic for future work. } \\subsection{Organization of the Paper} In \\S \\ref{sec:eom} we introduce the equations of motion, and in \\S \\ref{sec:pseudo} we present two pseudospectral simulations. One illustrates the formation and survival of a vortex in a short box, and the other illustrates the destruction of a vortex in a tall box. In \\S\\S \\ref{sec:lin}-\\ref{sec:nonlin} we develop a theory explaining this behavior. The reader who is satisfied by the qualitative description leading to equation (\\ref{eq:res}) may skip those two sections. The theory that we develop is indirectly related to the transient amplification scenario for the generation of turbulence. Even though hydrodynamical disks are linearly stable, linear perturbations can be transiently amplified before they decay, often by a large factor. It has been proposed that sufficiently amplified modes might couple nonlinearly, leading to turbulence \\citep[e.g.,][]{CZTL03,Yecko04,AMN05}. However, to make this proposal more concrete, one must work out how modes couple nonlinearly. In L07, we did that in two dimensions. We showed that the 2D nonlinear instability of equation (\\ref{eq:2dinst}) is a consequence of the coupling of an axisymmetric mode with a transiently amplified mode, which may be called a ``swinging mode'' because its phasefronts are swung around by the background shear. In \\S \\ref{sec:nonlin} we show that the 3D instability responsible for the destruction of vortices is a generalization of this 2D instability. It may be understood by examining the coupling of a 3D swinging mode with an axisymmetric one. 3D modes become increasingly unstable as $|k_z|$ decreases, and in the limit that $k_z\\rightarrow 0$, the 3D instability matches smoothly onto the 2D one. Thicker disks are more prone to 3D instability because they encompass smaller $|k_z|$. ", "conclusions": "\\label{sec:conc} Our main result follows from Figure \\ref{fig:marg}, which maps out the stability of a ``mother mode'' (i.e., a mode with wavevector $\\bar{k}\\bld{\\hat{x}}$ and amplitude $\\bar{\\omega}$) to nonaxisymmetric 3D perturbations. A mother mode is unstable provided that the $k_y$ and $k_z$ of the nonaxisymmetric perturbations satisfy both $|k_y|\\lesssim \\bar{k}\\bar{\\omega}/q$ and $|k_z|\\lesssim |k_y|$, dropping order-unity constants. Based on this result, we may understand the formation, survival, and destruction of vortices. Vortices form out of mother modes that are unstable to 2D ($k_z=0$) perturbations. Mother modes that are unstable to 2D modes but stable to 3D ($k_z\\ne 0$) ones, form into long-lived vortices. Mother modes that are unstable to both 2D and 3D modes are destroyed. Therefore a mother mode with given $\\bar{k}$ and $\\bar{\\omega}$ will form into a vortex if the disk has a sufficiently large circumferential extent and a sufficiently small scale height, i.e., if $r\\gtrsim \\bar{k}^{-1}q/\\bar{\\omega}$ and $h\\lesssim \\bar{k}^{-1}q/\\bar{\\omega}$, where $r$ is the distance to the center of the disk, and $h$ is the scale height. Alternatively, the mother mode will be destroyed in a turbulent-like state if both $r$ and $h$ are sufficiently large ($r\\gtrsim \\bar{k}^{-1}q/\\bar{\\omega}$ and $h\\gtrsim \\bar{k}^{-1}q/\\bar{\\omega}$). Our result has a number of astrophysical consequences. In protoplanetary disks that do not contain any vortices, solid particles drift inward. Gas disks orbit at sub-Keplerian speeds, $v_{\\rm gas}\\sim \\Omega r (1-\\eta)$, where $\\Omega r$ is the Keplerian speed and $\\eta\\sim (c_s/\\Omega r)^2$, with $c_s$ the sound speed. Since solid particles would orbit at the Keplerian speed in the absence of gas, the mismatch of speeds between solids and gas produces a drag on the solid particles, removing their angular momentum and causing them to fall into the star. For example, in the minimum mass solar nebula, meter-sized particles fall in from 1 AU in around a hundred years. This rapid infall presents a serious problem for theories of planet formation, since it is difficult to produce planets out of dust in under a hundred years. Vortices can solve this problem \\citep{BS95}. A vortex that has excess vorticity $-\\bar{\\omega}$ and radial width $1/\\bar{k}$ can halt the infall of particles provided that $\\bar{\\omega}/\\bar{k}\\gtrsim (\\Omega r)\\eta$, because the gas speed induced by such a vortex more than compensates for the sub-Keplerian speed induced by gas pressure.\\footnote{ We implicitly assume here that the stopping time of the particle due to gas drag is comparable to the orbital time, which is true for meter-sized particles at 1 AU in the minimum mass solar nebula. A more careful treatment shows that a vortex can stop a particle with stopping time $t_s$ provided that $\\bar{\\omega}/\\bar{k}\\gtrsim (\\Omega r)(\\Omega t_s)\\eta$ \\citep{Youdin08}. } Previous simulations implied that 3D vortices rapidly decay, and so cannot prevent the rapid infall of solid particles \\citep{BM05b,SSG06}. Our result shows that vortices can survive within disks, and so restores the viability of vortices as a solution to the infall problem. A more important---and more speculative---application of our result is to the transport of angular momentum within neutral accretion disks. In our simulation of a vortex in a tall box, we found that as the vortex decayed it transported angular momentum outward at a nearly constant rate for hundreds of orbital times. If decaying vortices transport a significant amount of angular momentum in disks, they would resolve one of the most important outstanding questions in astrophysics today: what causes hydrodynamical accretion disks to accrete? To make this speculation more concrete, one must understand the amplitude and duration of the ``turbulence'' that results from decaying vortices. This is a topic for future research. In this paper, we considered only the effects of rotation and shear on the stability of vortices, while we neglected the effect of vertical gravity. There has been a lot of research in the geophysical community on the dynamics of fluids in the presence of vertical gravity, since stably stratified fluids are very common on Earth---in the atmosphere, oceans, and lakes. In numerical and laboratory experiments of strongly stratified flows, thin horizontal ``pancake vortices'' often form, and fully developed turbulence is characterized by thin horizontal layers. \\citep[e.g., ][]{BBLC07}. Pancake vortices are stabilized by vertical gravity, in contrast to the vortices studied in this paper which are stabilized by rotation. Gravity inhibits vertical motions because of buoyancy: it costs gravitational energy for fluid to move vertically. The resulting quasi-two-dimensional flow can form into a vortex.\\footnote{\\cite{BC00} show that a vertically uniform vortex column in a stratified (and non-rotating and non-shearing) fluid suffers an instability (the ``zigzag instability'') that is characterized by a typical vertical lengthscale $\\lambda_z\\sim U/N$, where $U$ is the horizontal speed induced by the vortex, $N$ is the Brunt-V\\\"ais\\\"al\\\"a frequency, and the horizontal lengthscale of the vortex $L_h$ is assumed to be much greater that $\\lambda_z$ (hence the pancake structure). We may understand Billant \\& Chomaz's result in a crude fashion with an argument similar to that employed in the introduction to explain the destruction of rotation-stabilized vortices: since the frequency of buoyancy waves is $Nk_x/k_z$ (when $|k_x|\\ll |k_z|$), and since the frequency at which fluid circulates around a vortex is $U/L_h\\sim k_xU$, there is a resonance between these two frequencies for vertical lengthscale $1/k_z\\sim U/N$.} We may speculate that in an astrophysical disk vertical gravity provides an additional means to stabilize vortices, in addition to rotation. But to make this speculation concrete, the theory presented in this paper should be extended to include vertical gravity. We have not addressed in this paper the origin of the axisymmetric structure (the mother modes) that give rise to surviving or decaying vortices. One possibility is that decaying vortices can produce more axisymmetric structure, and therefore they can lead to self-sustaining turbulence. This seems to us unlikely. We have not seen evidence for it in our simulations, but this could be because of the modest resolution of our simulations. Other possibilities for the generation of axisymmetric structure include thermal instabilities, such as the baroclinic instability, or convection, or stirring by planets. This, too, is a topic for future research. \\appendix" }, "0710/0710.3059_arXiv.txt": { "abstract": " ", "introduction": " ", "conclusions": "\\label{sec:discussion} We considered the relation between the HMXB population and the SFH of the host galaxy. The number of HMXBs can be represented as a convolution (Eq. (4)) of the star formation history SFR(t) with the function $\\eta_{HMXB}(t)$ describing the dependence of the HMXB number on the time elapsed since the star formation event. Thus, the evolution of the HMXB population after the star formation event can be reconstructed by analyzing the distribution of HMXBs in stellar complexes with different SFHs. Using archival optical observations, we reconstructed the spatially resolved SFH in the SMC over the past $\\sim$100~Myr (Fig. 10). For this purpose, the observed color-magnitude diagrams of the stellar population were approximated by linear combinations of model isochrones. We analyzed the stability and errors of this method for reconstructing the recent SFH and showed that its accuracy is limited by the uncertainties in the currently available models for the evolution of massive stars. However, the systematic error introduced by this factor may be ignored, since the main source of uncertainty in the solution is the Poisson noise due to the relatively small number of HMXBs in the part of the SMC investigated by XMM-Newton. Using the derived SFHs and the spatial distribution of HMXBs in the SMC from Shtykovskiy and Gilfanov (2005b), we reconstructed the function $\\eta_{HMXB}(t)$ that describes the dependence of the HMXB number on the time elapsed since the star formation event (Fig. 12). We compared the derived dependence with the behavior of the SN II rate. The HMXB number reaches its maximum $\\sim$20--50~Myr after the star formation event, which is comparable to or exceeds the lifetime of a $8M_\\odot$ star. This is much later than the maximum of the SN II rate. In addition, note the shortage of the youngest systems. Observationally, this manifests itself in the absence (or an extremely small number) of HMXBs with black holes in the SMC. This behavior is related to the evolution of the companion star and the neutron star spin period and is consistent with the population synthesis model calculations (Popov et al. 1998). When these results are interpreted, it should be kept in mind that the function $\\eta_{HMXB}(t)$ depends on the luminosity threshold used to select the X-ray sources. In our analysis, we used a sample with a low luminosity threshold, L$_{min}\\sim 10^{34}$~erg/s. In such a sample, low-luminosity sources, mostly Be/X systems, mainly contribute to the number of sources, while the relative contribution from systems with black holes and/or O/B supergiants, which must constitute the majority of sources in the lower time bin in Fig. 12, is small. Therefore, the time dependence of the number of bright sources (e.g., $>10^{37}$~erg/s) will differ from that shown in Fig. 12. The HMXB formation efficiency in the SMC does not exceed the prediction of the N$_{HMXB}$--SFR calibration (Grimm et al. 2003). Their abnormal abundance compared to the predictions based on the emission in standard SFR indicators, such as the H$_{\\alpha}$ line, can result from a peculiarity of the SFH in the SMC." }, "0710/0710.4509_arXiv.txt": { "abstract": "We present a scheme to solve the nonlinear multigroup radiation diffusion (MGD) equations. The method is incorporated into a massively parallel, multidimensional, Eulerian radiation-hydrodynamic code with adaptive mesh refinement (AMR). The patch-based AMR algorithm refines in both space and time creating a hierarchy of levels, coarsest to finest. The physics modules are time-advanced using operator splitting. On each level, separate ``level-solve'' packages advance the modules. Our multigroup level-solve adapts an implicit procedure which leads to a two-step iterative scheme that alternates between elliptic solves for each group with intra-cell group coupling. For robustness, we introduce pseudo transient continuation ($\\ptc$). We analyze the magnitude of the $\\ptc$ parameter to ensure positivity of the resulting linear system, diagonal dominance and convergence of the two-step scheme. For AMR, a level defines a subdomain for refinement. For diffusive processes such as MGD, the refined level uses Dirichet boundary data at the coarse-fine interface and the data is derived from the coarse level solution. After advancing on the fine level, an additional procedure, the sync-solve (SS), is required in order to enforce conservation. The MGD SS reduces to an elliptic solve on a combined grid for a system of $G$ equations, where $G$ is the number of groups. We adapt the ``partial temperature'' scheme for the SS; hence, we reuse the infrastructure developed for scalar equations. Results are presented. We consider a multigroup test problem with a known analytic solution. We demonstrate utility of $\\ptc$ by running with increasingly larger timesteps. Lastly, we simulate the sudden release of energy $Y$ inside an Al sphere ($r = 15$ cm) suspended in air at STP\\@. For $Y = 11$ kT, we find that gray radiation diffusion and MGD produce similar results. However, if $Y = 1$ MT, the two packages yield different results. Our large $Y$ simulation contradicts a long-standing theory and demonstrates the inadequacy of gray diffusion. ", "introduction": "\\label{intro} This paper describes a numerical method to solve the radiation multigroup diffusion (MGD) equations. Two themes are presented. One is the scheme itself. We add Pseudo Transient Continuation $(\\ptc)$ to the familiar ``fully implicit'' method of Axelrod et al \\cite{AxDuRh}. The second theme is code-specific. Our MGD solver is embedded in a multidimensional, massively parallel, Eulerian radiation-hydrodynamic code, which has patch-based, time-and-space Adaptive Mesh Refinement (AMR) capability. Our code's AMR framework stems from the Berger and Oliger idea \\cite{BeOl} developed for hyperbolic, compressible hydrodynamic schemes. The idea was expanded by Almgren et al \\cite{Alm} and applied to the type of elliptic solvers required for the incompressible equations of Navier-Stokes. Howell and Greenough \\cite{HowGre} applied the Almgren et al framework to the scalar, parabolic ``gray'' radiation diffusion equation, thereby creating the start of our radiation-hydrodynamic code. The AMR framework works as follows. A domain, referred to as the ``coarse'' or L0 level, is discretized using a uniform, coarse spatial mesh size $h_c$.\\footnote{In multiple dimensions, coordinates have their own mesh spacing.} After advancing with a timestep $\\dtc$, the result is scanned for possible improvement. One may refine subregions containing a chosen material, at material interface(s), or at shocks, etc. Whatever refinement criteria are used, after the subdomains are identified, specific routines define a collection of ``patches,'' which cover the subdomains. In two dimensions, the patches are unions of rectangles; in 3D, they are unions of hexahedra. The patches need not be connected, but they must be contained within the coarse level. The patches denote the ``fine'' or L1 level and are discretized with a uniform, spatial mesh size $h_f$. A typical refinement ratio $h_c/h_f$ equals two, but higher multiples of two are also allowed. Because the original framework was designed for temporally explicit hyperbolic schemes, $\\dtc$ is restricted by a CFL condition. This implies a similar restriction for the L1 level timestep $\\dtf$. For the case, $h_c/h_f = 2$, level L1 time-advances twice using $\\dtf = \\dtc / 2$. Boundary conditions for level L1 are supplied as follows. Wherever level L1 extends to the physical boundary, the level uses the conditions prescribed by the problem. Portions of level L1's boundary which lie inside the physical domain have conditions prescribed by time and space interpolated data obtained from the L0 solution. For diffusion equations, these conditions are of Dirichlet type. The numerical solution consists of both coarse and fine grid results. Unfortunately, as it stands, the composite solution does not guarantee conservative fluxes across the level boundaries. To maintain conservation, a separate procedure, dubbed a sync-solve (SS) is required. The SS reduces to an elliptic unstructured grid solve on the composite grid of L0 and L1 levels. The AMR procedure may be recursive. That is, a level L1 grid may generate its own subdomain for refinement, i.e., a level L2. In that case, one SS couples results from levels L1 and L2. Once the levels advance to the L0 level time, a SS coupling all three levels ensues. For the multigroup equations, the SS requires an unstructured grid solve for a coupled system of reaction-diffusion equations. Our scheme for a multigroup SS is an important theme of this paper. The MGD equations stem from a discretization of the multifrequency radiation diffusion equations. The latter is an approximation to the equations of radiation transfer, obtained by assuming the matter to be optically thick, which suppresses the directional dependence of the radiation intensity. Details of the derivation may be found in various sources: Mihalas and Mihalas \\cite{MM2}, Zel'dovich and Raizer \\cite{ZelRai}, Pomraning \\cite{Pom}. The gray radiation diffusion equation is a simplification of the MGD equations. It is essentially a one-group equation and is derived by integrating over all frequencies. Surprisingly, it gives very good results in many cases. However, it clearly cannot display frequency-dependent effects. When those are important, it gives incorrect results. Unfortunately, unless one solves a problem with both gray and MGD, one never knows when the former is adequate. We now summarize the paper. Our MGD scheme consists of two parts. Sections~\\ref{levelsolve} and \\ref{MGanlsys} develop the level-solve algorithm, which is applied on each level. Section~\\ref{levelsolve} develops the equations, the discretization, and our $\\ptc$ scheme. Section~\\ref{MGanlsys} proves three lemmas which determine the initial magnitude of the $\\ptc$ parameter $\\gs$. % Our philosophy for $\\gs$ is as follows. The result of the level solve is the time-advanced radiation group energy density, which physics dictates to be nonnegative. Zeroing anomalously negative values is not an option since they are the correct conservative solution to the linear system that stems from the discretization of the system. Thus, the unphysical result nonetheless conserves energy. The difficulty is avoided if in the original formulation of the linear system $ A x = b$, $A$ is an M-matrix and the right-hand-side (RS) is nonnegative. Since we solve $ A x = b$ using an iterative scheme, the magnitude of $\\gs$ is determined to ensure $b \\ge 0$, a diagonally dominant $A$, and that the iterations converge. To a large extent, we are guided by Pert~\\cite{Pert}, who discusses how and why the solution to a discretization of an equation may be unacceptable from a physical standpoint. For a first reading, section~\\ref{MGanlsys} may be skipped; the analysis of the required magnitude of $\\gs$ is not needed for the subsequent sections. We note that $\\ptc$ is widely used to solve nonlinear systems of equations. It is closely related to the Inexact Newton Backtracking Method by Shahid et al \\cite{ShTuWa}. When applying $\\ptc$ to a Newton solver, the basic idea is to limit the change to the iterates when one is far from the root but not restrict the change as one approaches the root. With $\\ptc$, limiting is done by the magnitude of the pseudo-timestep. Kelley and Keyes \\cite{KelKey} put $\\ptc$ on a solid analytic framework by examining the three regimes of $\\ptc$: small, medium, and large pseudo-timesteps. In the last regime, $\\ptc$ recovers Newton's second order of convergence. Our $\\ptc$ implementation differs from the norm. Standard applications typically detect when a problem is ``hard'' and then reduce the timestep or some other parameter by an arbitrary amount. However, this method will not work for us because our solver is embedded in a time-dependent multiphysics code with separate modules for compressible gasdynamics, heat conduction, radiation transport. Our MGD solver is called numerous times during the course of a simulation. (If running, with AMR, multiple times per physical time advance.) Although the physical $\\dt$ is controlled by various means, and depending on the problem can vary many orders of magnitude, we require a MGD solver that works under all conditions. Our $\\ptc$ approach is similar to the one of Shestakov et al \\cite{ShHoGr}. We set the initial magnitude of the $\\ptc$ parameter to ensure that for the first step, our iteration scheme converges and that the result is physical. We note that our usage of $\\ptc$ is nearly equivalent to having the MGD module time-advance not in a single (physical) step $\\dt$, but in smaller time increments until the desired time $t^0 + \\dt$ is reached. Some colleagues refer to the process as ``sub-cycling'' the radiation module. It is easy to show that the lemmas of Sec.~\\ref{MGanlsys} still apply for sub-cycling. Section~\\ref{mgamr} describes the second part of our solver, viz., the sync-solve. Section~\\ref{rapAMR} contains results. Three problems are presented. The first, in Sec.~\\ref{linwin}, displays the accuracy of the method and its convergence properties: first order in time and second order in space. Section~\\ref{PTCrobust} demonstrates the utility afforded by $\\ptc$. For hard problems, it accelerates convergence; for very hard problems, $\\ptc$ is indispensable. Section~\\ref{hotball} models the explosive expansion of a hot metal sphere suspended in cold air. The simulation couples all of the code's physics modules. The problem is an ideal candidate for AMR since effects propagate a large distance away from the source, yet in early times, resolution is needed only near the sphere. The problem also demonstrates the necessity of multigroup diffusion. We find that if the sphere's energy is very high, gray diffusion gives the wrong answer. For a 1 MT energy source, our MGD simulation contradicts results of Brode \\cite{Bro}, who used gray diffusion. Section~\\ref{conclusion} contains concluding remarks. There are three appendices. Appendix~\\ref{table} gives a table of exact values for the test problem described in section~\\ref{linwin}. Appendix~\\ref{apb} discusses situations that may complicate attaining a diagonally dominant matrix when discretizing the multigroup system. Appendix~\\ref{apc} presents a spatial convergence analysis of the multigroup system when running in ``production'' mode, that is, with a dominant flux limiter and with AMR. ", "conclusions": "\\label{conclusion} We have described a numerical scheme to solve the radiation multigroup diffusion equations. The scheme is implemented in a radiation-hydrodynamic code with the patch-based AMR methodology, originally proposed by Berger and Oliger \\cite{BeOl} for hyperbolic partial differential equations. Our scheme consists of two parts. The first, described in Sections~\\ref{levelsolve} and \\ref{MGanlsys}, is applied on a {\\em level\\/} of the AMR grid layout and may be adapted to any code. This part consists of adding $\\ptc$ to the ``fully-implicit'' iterative scheme of Axelrod et al \\cite{AxDuRh}. $\\ptc$ brings an extra degree of robustness and enhances convergence of the Axelrod scheme. We have developed lemmas that determine the minimum magnitude for the $\\ptc$ parameter $\\tau$ to ensure that the iterations converge and the result is physically meaningful. The appropriate magnitude depends on the problem. Our implementation of $\\ptc$ is not optimal---at least for our AMR code architecture. In our code, for each AMR level, we compute a {\\em single\\/} scalar parameter $\\tau$. However, the levels consist of a collection of grids (rectangles in 2D) that need not be connected. If the grids are not connected, they form independent problems. Hence, it would be more efficient to use different $\\tau$ for disconnected grids. The second part of our scheme, the sync-solve (SS), addresses a specific need of our code, viz., the requirement of having an energy-conserving result on the composite grid of multiple AMR levels. For the multigroup equations, this part reduces to a coupled system of elliptic equations on the unstructured grid combining all levels. Since the SS is intended to be a small correction to the result of the level solves, we adapted the key element of the ``partial temperature'' scheme of Lund and Wilson \\cite{LunWil}. This allowed reducing the multigroup SS to a collection of scalar SS's. We were then able to reuse existing software.% This paper included simulations of three problems. The first two are idealized tests of only the multigroup module. The third is a ``real'' problem, which uses the full capability of the code: AMR, multiple materials, etc. The first problem was chosen because of its non-triviality and the availability of analytic results with which to compare. We obtained excellent agreement and verified the convergence properties of the scheme. The second problem illustrated the benefits brought by $\\ptc$. We compared the conventional scheme of Axelrod et al \\cite{AxDuRh} with our $\\ptc$-modified version. For hard problems, $\\ptc$ either decreased run times or ensured convergence in regimes where the conventional scheme diverged. The third problem showed that our multigroup module has been fully integrated into the code and has already extended the scientific frontier. For a high yield air burst at STP, we found that gray diffusion gives an incorrect result during the radiation-dominated regime because gray fails to capture the frequency-dependent effects of the air opacity." }, "0710/0710.2213_arXiv.txt": { "abstract": "We investigate gravitino dark matter scenarios in which the primordial $\\Lisix$ production is catalyzed by bound-state formation of long-lived negatively charged particles $\\champ$ with $\\Hefour$. In the constrained minimal supersymmetric Standard Model (CMSSM) with the stau $\\stau^-$ as the $\\champ$, the observationally inferred bound on the primordial $\\Lisix$ abundance allows us to derive a rigid lower limit on the gaugino mass parameter for a standard cosmological history. This limit can have severe implications for supersymmetry searches at the Large Hadron Collider and for the reheating temperature after inflation. ", "introduction": "\\label{sec:introduction} Big Bang Nucleosynthesis (BBN) is a powerful tool to test physics beyond the Standard Mo\\-del. Recently, it has been realized that the presence of heavy long-lived negatively charged particles $\\champ$ can have a substantial impact on the primordial light element abundances via bound-state formation% ~\\cite{Pospelov:2006sc,Kohri:2006cn,Kaplinghat:2006qr,Cyburt:2006uv,Hamaguchi:2007mp,Bird:2007ge,Kawasaki:2007xb,Jittoh:2007fr,Jedamzik:2007cp}. In particular, when $\\champ$ and $\\Hefour$ form Coulomb bound states, $(\\Hefour\\champ)$, too much $\\Lisix$ can be produced via the catalyzed BBN (CBBN) reaction~\\cite{Pospelov:2006sc} \\begin{align} \\label{eq:CBBN-reaction} (\\Hefour\\champ)+\\mathrm{D} \\rightarrow \\Lisix + \\champ . \\end{align} The formation of $(\\Hefour\\champ)$ and hence the CBBN production of $\\Lisix$ becomes efficient at temperatures $T \\sim 10\\ \\keV$, i.e., at cosmic times $t> 10^3\\ \\seconds$ at which standard BBN (SBBN) processes are already frozen out. The observationally inferred bound on the primordial $\\Lisix$ abundance then restricts severely the $\\champ$ abundance at such times. A long-lived $\\champ$ may be realized if the gravitino is the lightest supersymmetric particle (LSP). In particular, it is reasonable to consider gravitino LSP scenarios within the constrained minimal supersymmetric Standard Model (CMSSM)~\\cite{Ellis:2003dn,Cerdeno:2005eu,Jedamzik:2005dh,Cyburt:2006uv,Pradler:2006hh} in which the gaugino masses, the scalar masses, and the trilinear scalar couplings are parameterized by their respective universal values $\\monetwo$, $\\mzero$, and $A_0$ at the scale of grand unification $\\mgut \\simeq 2\\times 10^{16}\\ \\GeV$. Within this framework, the lighter stau $\\stau$ is the lightest Standard Model superpartner in a large region of the parameter space and thus a well-motivated candidate for the next-to-lightest supersymmetric particle (NLSP). Since its couplings to the gravitino LSP are suppressed by the (reduced) Planck scale, $\\MP=2.4\\times 10^{18}\\,\\GeV$, the stau will typically be long-lived for conserved $\\mathrm{R}$-parity% \\footnote{For the case of broken R-parity, see, e.g.~\\cite{Buchmuller:2007ui}} and thus $\\stau^-$ can play the role of $\\champ$. In scenarios with conserved $\\mathrm{R}$-parity, the gravitino LSP is stable and a promisig dark matter candidate. Gravitinos can be produced efficiently in thermal scattering of particles in the primordial plasma. If the Universe, after inflation, enters the radiation dominated epoch with a high reheating temperature $\\TR$, the resulting gravitino density $\\Omegatp$ will contribute substantially to the dark matter density $\\Omega_{\\mathrm{dm}}$~\\cite{Bolz:2000fu,Pradler:2006qh,Rychkov:2007uq}. In this work we calculate the amount of $\\Lisix$ produced in~(\\ref{eq:CBBN-reaction}) by following the treatment of Ref.~\\cite{Takayama:2007du}. In particluar, we employ a recent state-of-the-art result for the CBBN reaction cross reaction~\\cite{Hamaguchi:2007mp}. The obtained upper limit on the $\\champ$ abundance from possible $\\Lisix$ overproduction vanishes for sufficiently short $\\tau_{\\champ}$. This allows us to extract a lower limit on the universal gaugino mass parameter $\\monetwo$ within minimal supergravity scenarios where the gravitino is the LSP and the $\\champ$ is the $\\stau^-$ NLSP.\\footnote{In this work we assume a standard cosmological history with a reheating temperature $\\TR$ that exceeds the freeze-out temperature $\\Tf$ of the $\\stau$ NLSP.} This limit leads directly to an upper bound on $\\TR$ since $\\Omegatp$ cannot exceed the observed dark matter density. The bounds on $\\monetwo$ and $\\TR$ derived below depend on the gravitino mass but are independent of the CMSSM parameters. Before proceeding, let us comment on the present status of BBN constraints on gravitino dark matter scenarios with a long-lived charged slepton NLSP. In a recent ambitious study~\\cite{Jedamzik:2007cp} it is argued that bound-state formation of $\\champ$ with protons at $T \\sim 1\\ \\keV$ might well reprocess large fractions of the previously synthesized $\\Lisix$. This seems to relax the bound on the $\\champ$ abundance for $\\tau_{\\champ}> 10^6~\\seconds$. However, at present, the uncertainties in the relevant nuclear reaction rates in~\\cite{Jedamzik:2007cp} make it difficult to decide whether a new cosmologically allowed region will open up. In this work we assume that this is not the case, in particular, since the $^3$He/D constraint on electromagnetic energy release~\\cite{Sigl:1995kk} becomes severe in this region and excludes stau lifetimes $\\tau_{\\stau}\\gtrsim10^6~\\seconds$~\\cite{Cerdeno:2005eu,Cyburt:2006uv,Kawasaki:2007xb,Jedamzik:2007cp}. Then only the constraint from hadronic energy release on D~\\cite{Kawasaki:2004qu,Feng:2004mt,Cerdeno:2005eu,Jedamzik:2006xz,Steffen:2006hw} can be slightly more severe than the one from catalyzed $^6$Li production~\\cite{Cyburt:2006uv,Steffen:2006wx,Pradler:2006hh,Kawasaki:2007xb}. We neglect the D constraint in this work since it can only tighten the bounds on $\\monetwo$ and $\\TR$ as can be seen, e.g., in Figs.~4~(b--d) and~5 of Ref.~\\cite{Pradler:2006hh}. For deriving conservative bounds on $\\monetwo$ and $\\TR$, it is thus sufficient to consider the CBBN reaction~(\\ref{eq:CBBN-reaction}) exclusively. ", "conclusions": "\\label{sec:conclusion} We have considered the catalysis of $\\Lisix$ production in CMSSM scenarios with the gravitino LSP and the stau NLSP. Within a standard cosmological history, the calculated $\\Lisix$ abundance drops below the observational limit on primordial $\\Lisix$ for $\\tau_{\\stau} \\lesssim 5\\times 10^3\\,\\seconds$. Taken at face value, we find that this constraint translates into a lower limit $\\monetwo \\ge 0.9\\, \\TeV ( \\mgr / 10\\, \\GeV )^{2/5}$ in the entire natural region of the CMSSM parameter space. This implies a conservative upper bound $\\TR\\lesssim 4.9\\times 10^7\\GeV( \\mgr / 10\\, \\GeV )^{1/5}$. The bounds on $\\monetwo$ and $\\TR$ not only confirm our previous findings~\\cite{Pradler:2006hh} but are also independent of the particular values of the CMSSM parameters for the considered $\\stau$ NLSP abundances." }, "0710/0710.2814_arXiv.txt": { "abstract": "CO isotopes are able to probe the different components in protostellar clouds. These components, core, envelope and outflow have distinct physical conditions and sometimes more than one component contributes to the observed line profile. In this study we determine how CO isotope abundances are altered by the physical conditions in the different components. We use a 3D molecular line transport code to simulate the emission of four CO isotopomers, $^{12}$CO $J=2\\rightarrow1$, $^{13}$CO $J=2\\rightarrow1$, C$^{18}$O $J=2\\rightarrow1$ and C$^{17}$O $J=2\\rightarrow1$ from the Class 0/1 object L483, which contains a cold quiescent core, an infalling envelope and a clear outflow. Our models replicate JCMT (James Clerk Maxwell Telescope) line observations with the inclusion of freeze-out, a density profile and infall. Our model profiles of $^{12}$CO and $^{13}$CO have a large linewidth due to a high velocity jet. These profiles replicate the process of more abundant material being susceptible to a jet. C$^{18}$O and C$^{17}$O do not display such a large linewidth as they trace denser quiescent material deep in the cloud. ", "introduction": "Molecules, particularly CO, are used as tracers of H$_2$ and thus of gas density in cold dark clouds. CO is highly abundant with a low critical density and typically exhibits strong optical depth effects in cold dark molecular clouds. The four most common CO isotopes differ in abundance by as much as three orders of magnitude. Thus they become optically thick at different column densities. Taken together, observations of CO isotopes can trace the gas density in all the main components of cold, dark clouds: the intermediate optical depth envelope, the high optical depth core, and the optically thin bipolar outflows. However, we know from observational and theoretical studies that the abundance of CO depends on conditions in the clouds such as, shock heating \\citep{van_dishoeck_95,nisini_07}, U.V. excitation \\citep*{goldsmith_07}, freeze-out \\citep*{lee_04} and varies from place to place. Therefore we cannot use CO as an H$_2$ tracer without understanding its chemical variation. To better understand CO variations in cold dark clouds, we observed and modeled one particular cloud, Lynds 483. We chose L483 as a prototype for a study of CO abundances because it is a well studied nearby ($\\sim 200~{\\rm pc}$) molecular cloud. It contains an IRAS source 18148-0440 that is in transition between a Class 0 and Class 1 object \\citep{tafalla.et.al00}. It exhibits an infalling envelope \\citep{park.et.al99,park.et.al00,tafalla.et.al00} and a slow bipolar molecular outflow \\citep{fuller.et.al95,buckle.et.al99} yet the core and envelope are still cold and dense \\citep{ladd.et.91,fuller.et.92,fuller&wootten00}. Thus many of the physical properties and kinematic features that are present in either less or more evolved clouds are all present in L483. Our aim is to combine these components into a single model for L483 to strongly constrain the structure and dynamics of the system and hence then to infer the CO abundance throughout the cloud. \\begin{figure*} \\centering \\includegraphics[width=180mm]{fig1} \\caption{An example line profile of each of the four isotopes used in this study where the offsets with respect to IRAS 18140-0440 are indicated in the top right corner of each panel. The $^{12}$CO y-axis is in units of $T_{r}$ and $^{13}$CO, C$^{18}$O and C$^{17}$O are in units of $T_{mb}$ } \\label{overview} \\end{figure*} We obtained emission line profile data from L483 for transitions of the four most common CO isotopes (section ~\\ref{obs}). In addition, we used an archival dust continuum emission map of L483 as an unbiased mass tracer to be compared with the CO data. We used a radiative transfer model to analyse the data and to produce synthetic spectra to be compared in detail with individual observations (section ~\\ref{models}). Estimates for density, temperature and a physical model of L483 were constrained with help from the observational literature. The abundances of the CO isotopes were then varied in the model to give a good match with the observed line profiles. This yields the abundance variation throughout L483 as well as self-consistent temperatures, densities and velocities and abundance ratios. ", "conclusions": "We modelled L483 in four isotopomers of CO in order of decreasing abundance $^{12}$CO, $^{13}$CO, C$^{18}$O and C$^{17}$O. Each species delineates a different region and therefore we get a clearer picture of cloud dynamics. C$^{18}$O and C$^{17}$O are optically thin and are important tracers of the denser regions of star forming clouds. Optically thick species such as $^{12}$CO and $^{13}$CO are more abundant and more susceptible to the jet motion. They trace regions farther out from the centre of the cloud and have line profiles showing a large line width consistent with material exposed to a high velocity jet. Using a radiative transfer code and a dynamical model for L483 we were able to self-consistently calculate for the first time the abundance of the CO isotopes in the different regions of such a cloud. Our principal finding is that the CO line profiles in L483 are well fitted with a self-consistent envelope plus boundary layer model and that the CO abundances increase substantially in this boundary layer. The most likely reason for this is that molecular ices on dust grains are heated and released back into the gas phase in the boundary layer. A constant abundance model was found to overestimate the abundance towards the centre of the cloud and only freeze-out of material towards the centre was able to produce modeled profiles consistent with observations. Our tanh geometry is chosen because it matches the observed morphology seen in other protostellar outflows \\citep{tafalla.et.al00}. Other geometries are possible for the outflow e.g. conical, cylindrical outflow but it is unlikely that it would have a substantial effect on our results because such a detailed treatment of the boundary layer components is difficult to achieve until there is sub-arcsecond resolution resolution. We emphasize that our results provide an abundance enhancement measurement rather than proving an exact mechanism by which the CO is enhanced, e.g. chemical reactions or dissociation. A more detailed treatment would involve a full dark cloud and gas-grain chemistry whilst accounting for localized shock heating. The most enhanced species in our study, by a factor of $\\sim 30$, is the $^{13}$CO material. The exothermic reaction leading to the creation of $^{13}$CO \\citep{duley_williams} is shown below \\begin{equation} \\; \\; \\; \\; \\; ^{13}{\\rm C}^{+} + ^{12}{\\rm CO} \\rightleftharpoons ^{12}{\\rm C}^{+} + ^{13}{\\rm CO} + \\Delta{\\rm E} \\label{13co_creation} \\end{equation} where the zero-point energy difference $\\Delta{\\rm E}$ is equivalent to a temperature $\\Delta{\\rm E}/{\\rm k}$ of 35~K. This mechanism may be the source for the enhanced abundance observed from our modeling. The enhanced abundance seen in the boundary layer effect may also be detectable in other molecules. \\citet{park.et.al00} used interferometric observations of HCO$^{+}$ and observed anti-infall profiles close to the centre of the cloud. They concluded the HCO$^{+}$ was tracing the outlying regions of the outflow, i.e. a region between the envelope and the jet. The reason the HCO$^{+}$ emission is predominately seen here rather than in the more extensive envelope is also likely due to an enhancement of HCO$^{+}$ caused by the shock-heated release of icy grain mantles followed by chemical reaction. CO and H$_2$O are liberated into the gas phase and the shock-induced radiation field then can photodissociate CO to C$^+$. This then reacts with the H$_2$O to form HCO$^{+}$. Such a model was successfully used to explain the enhancement of HCO$^{+}$ commonly seen at the bases of molecular outflows \\citep{rawlings.et.al00,rawlings.et.al04}. The results in this paper demonstrate that a combination of datasets with several lines and transitions coupled with a 3D molecular line transport code is a powerful way to determine the properties of dense star forming cores." }, "0710/0710.5119_arXiv.txt": { "abstract": "{The excess above 1 GeV in the energy spectrum of the diffuse Galactic gamma radiation, measured with the EGRET experiment, can be interpreted as the annihilation of Dark Matter (DM) particles. The DM is distributed in a halo around the Milky Way. Considering the directionality of the gamma ray flux it is possible to determine the halo profile. The DM within the halo has a smooth and a clumpy component. These components can have different profiles as suggested by N-body simulations and the data is indeed compatible with a NFW profile for the diffuse component and a cored profile for the clumpy component. These DM clumps can be partly destroyed by tidal forces from interactions with stars and the gravitational potential of the Galactic disc. This effect mainly decreases the annihilation signal from the Galactic centre (GC). In this paper constraints on the different profiles and the survival probability of the clumps are discussed. \\PACS{ {95.35.+d}{Dark Matter} \\and {98.35.Gi}{Galactic halo} } % } % ", "introduction": "\\label{seq:intro} From WMAP measurements of the temperature aniso- tropies in the Cosmic Microwave Background (CMB) in combination with data on the Hubble expansion and the density fluctuation in the Universe \\cite{RefSpergel} we gather that Cold Dark Matter (CDM) makes up 23\\% of the energy of the Universe. The nature of the Dark Matter (DM) is unknown, but one of the most promising particles is the \"weakly interacting massive particle\" (WIMP). Assuming that WIMPS are Majorana particles they can annihilate each other and produce a large amount of secondary particles. For the determination of the density distribution, the so-called DM halo profile of the WIMP particles, the gamma radia\\-tion is most important because it is not influenced by the magnetic field of the galaxy and points back directly to its source.\\\\ The observation of the diffuse Galactic gamma radiation of the Milky Way with EGRET showed an excess above 1 GeV in the photon energy spectrum. This excess is different for various sky directions and can be interpreted as a WIMP annihilation signal \\cite{RefSander}. Therefore it can be used to determine the DM halo profile.\\\\ In section \\ref{seq:halo}, we will describe the mechanism of the determination of the DM halo profile from the EGRET excess and explain how to calculate the DM annihilation flux of gamma rays. Then, after differentiating between diffuse and clumpy DM, we will introduce a survival probability for DM clumps as well as a ringlike substructure of DM within the Galactic disc. ", "conclusions": "CDM probably consists of WIMPS which are heavy and slow particles. If these particles are Majorana particles they can annihilate each other and produce Galactic gamma radiation which can be used to determine the density profile of the DM. In this analysis the directionality of the DM annihilation flux measured with EGRET was used to find a possible DM halo profile. After dividing the Galactic gamma ray flux into 4 latitude and 45 longitude regions the background and the DM annihilation signal were fitted to the data for each of the 180 bins. The DM annihilation flux is dominated by the annihilation flux of the clumpy DM which is proportional to $\\rho_{\\chi, clump}$, not $\\rho^2_{\\chi, clump}$. While the diffuse DM component has a cuspy NFW profile a shallower cored distribution was obtained for the clumpy component. The DM annihilation flux is dominated by the clumpy DM component, but the clumpy component yields a mass below the required mass $> 10^{12}$ solar masses \\cite{RefBattaglia}. However, if combined with the diffuse cuspy NFW profile, both the EGRET data and the mass constraint can be fulfilled. In order to take the tidal disruption of DM clumps in the vicinity of stars into account a survival probability for clumps was introduced. Most of the clumps are expected to be destroyed near the Galactic centre, although a steep cusp may survive. The strong signal observed from the Galactic centre yielded a survival probability at the centre of $P(0)=0.7$. This means that the DM clumps are not completely destroyed, which is in good agreement with more detailed calculations in Ref. \\cite{RefDokuchaev2}. In addition to the DM halo profiles two ringlike substructure were required at radii of 4 and 13 kpc. The halo and ring parameters were obtained by minimizing a $\\chi^2$ function comparing the flux of the excess from the various sky directions with the line-of-sight integral in the halo. Figure \\ref{fig:longitudes} shows that the halo model fits the measured data very well.\\\\ In summary, the EGRET excess of diffuse Galactic gamma rays is in good agreement with the expectations of a cored clumpy halo component plus a cuspy diffuse one. The ringlike substructure, expected from the tidal disruption of the nearby Canis Major dwarf galaxy, is clearly seen and its heavy mass above $10^{10}$ solar masses as obtained from the EGRET data, has been recently confirmed by the reduced gas flaring in this region." }, "0710/0710.5269_arXiv.txt": { "abstract": "We investigate the full $5D$ dynamics of general braneworld models. Without making any further assumptions we show that cyclic behavior can arise naturally in a fraction of physically accepted solutions. The model does not require brane collisions, which in the stationary case remain fixed, and cyclicity takes place on the branes. We indicate that the cosmological constants play the central role for the realization of cyclic solutions and we show that its extremely small value on the observable universe makes the period of the cycles and the maximum scale factor astronomically large. ", "introduction": "The last decade proves to be really exciting for cosmology. Observational data indicated, among other very interesting results, that the expansion of the universe is accelerated \\cite{observ}. At the same time the braneworld scenario appeared in the literature \\cite{Horawa,RS99}. Though the exciting idea that we live in a fundamentally higher-dimensional spacetime which is greatly curved by vacuum energy was older \\cite{Rubakov83}, the new class of ``warped\" geometries offered a simple way of localizing the low energy gravitons on the brane. In this novel background the old idea of a cyclic Universe was reheated. Started as ekpyrotic \\cite{Khoury.best,Khoury.rebound}, enriched to ekpyrotic/cyclic \\cite{ekpyrotic1,Khoury.rebound,Turok.simplified,Turok.sing,ekpyrotic.pert,cyclic.clifton,chargeBH} and recently to new ekpyrotic \\cite{ekpyrotic3,ekpyrotic4,ekpyrotic2,ekpyrotic.fields}, the new paradigm tries to be established as an alternative to standard cosmology. According to its basic contents, our universe experiences an infinite or extremely large number of cycles, each one consisting of a hot bang phase, a phase of accelerated expansion, a phase of slow-ekpyrotic contraction and a bounce-bang that triggers the next cycle. Starting with a simplified notional framework (infinite and not ``created\" time) cyclic cosmology have many advantages. It successfully faces the homogeneity, isotropy, topological and flatness problems, it handles the issue of initial conditions, it incorporates the dark energy and transforms it to an important factor, and it provides the mechanism of the generation of cosmic perturbations and of structure formation. However, there are some key issues that do not have a consistent and efficient approach so far, despite the great progress. These are the settlement of the singularity, although temperature and density remain finite, the entropy evolution, and the fate of the perturbations through the bounce. Through this research, cyclic scenarios have become more complicated, by the insertion of more complex potentials, of more branes \\cite{Khoury.best}, of the mechanism of ghost condensation \\cite{ghcond,ekpyrotic3}, of more scalar fields \\cite{ekpyrotic.fields} and of procedures which cancel the tachyonic instabilities \\cite{ekpyrotic4}. Most of the works on cyclic cosmology involve, initially or at some stage, the transition to effective $4D$ equations. However, as it was mentioned in \\cite{Linde01,4Dbreakdown}, such a procedure does not lead to reliable results since one cannot return to the $5D$ description self-consistently. Furthermore, the old $4D$-singularity problem (of both Big Bang and traditional cyclic universes), has been replaced by a new one (equally annoying) concerning the singularity of extra dimension(s). This later case is accompanied by the brane collision phenomenon, which seems to be a basic constituent of the ekpyrotic scenario. In this work we desire to investigate the full $5D$ dynamics of general braneworld models and examine if a cyclic behavior is possible. This is an essential procedure in order to consistently confront the arguments of the authors of \\cite{Linde01}, which claim that cyclic behavior cannot arise from a complete $5D$ description, and our study must not include any additional assumptions or fine tunings in order to remain general and therefore convincing. Secondly, we are interested to explore if a cyclic behavior of $5D$ dynamics is necessarily related to brane collisions. This work is organized as follows: In section \\ref{model} we present the $5D$ braneworld model and we derive the equations of motion. In section \\ref{analyt} we provide analytical solutions for two simplified stationary solution subclasses, while in \\ref{numer} we investigate numerically the full stationary dynamics. Finally, in section \\ref{discussion} we discuss the physical implications of our analysis and we summarize the obtained results. ", "conclusions": "\\label{discussion} In the aforementioned analysis we considered general braneworld models characterized by the action (\\ref{action}), the conformal metric (\\ref{metric}), and the general potentials (\\ref{bulkpot}) and (\\ref{branepot}). Performing both analytical and numerical calculations we showed that the full $5D$ dynamics allows for stationary solutions corresponding to oscillatory scale factor of the physical brane and therefore to cyclic universes. In statistical terms cyclicity corresponds to $\\approx 4\\%$ of the physical solutions. Our investigation is completely $5D$, cyclic behavior arises naturally and is induced on the brane by the full dynamics, and it is not a result of a modified $4D$ dynamics, with fine-tuned parameters or specific assumptions in the Friedmann equation. Furthermore, we do not use an explicit brane state equation, considering just the bulk scalar field (the decays and interactions of which will eventually fill the physical brane with the conventional content \\cite{apostol}). As we mentioned in the introduction this full $5D$ approach is necessary in order to confront the arguments of the authors of \\cite{Linde01}. Indeed, their allegations that one cannot transit to an effective $4D$ theory (integrating the action over $y$), solve the equations there and then return naively to the $5D$ description (adding time-dependence by hand), are correct. Doing so, the results are not self-consistent (especially the boundary conditions are not satisfied) and the authors of \\cite{Linde01} use this fact as a central argument against the cyclic scenario. However, our consistent $5D$ analysis reveals that cyclic behavior is possible. Another important feature of the present study is that cyclic universes do not require brane collisions. Thus, we avoid the known problems concerning such a description, which force ekpyrotic model to successively more complicated versions. On the contrary, the branes do not move at all and the system is stable (stationary solutions are a stable fixed point \\cite{brcod,Tetradis01}). Furthermore, in our model, expansion and contraction take place in the 3+1 branes, and in all 3+1 slices in general, while the fifth dimension remains unaffected. The $4$ spacial dimensions shrink periodically to an $1D$ string and re-expand. This is in a radical contrast with the cyclic models with extra dimensions, where the extra dimension is the one that gets contracted (the fifth in \\cite{ekpyrotic1} or the eleventh in \\cite{Khoury.rebound}). Cyclicity seems to re-obtain its ``physical\" meaning. Our $5D$ investigation is general and does not involve extra assumptions, fine-tunings or specific potential forms. We result to periodic, cyclic, homogenous and isotropic universes, where the scale factor changes smoothly from expanding to contracting. An observer on the physical brane feels successively accelerated expansion, decelerated expansion, turnabout, accelerated contraction, decelerated contraction, bounce e.t.c, and a promising signature of the cyclic behavior would be the measure of the varying rate of the Hubble constant. The cycles period, given in (\\ref{period}), can be arbitrary, depending on $B(0)$, i.e. on the value of the warp factor on the physical brane ($\\theta$ is bounded from above and therefore cannot act as a period-decreasing factor). A very interesting conclusion comes from the insertion of observational results in our model, which was not made above in order to remain as general as possible. Explicitly, if we use the fact that the cosmological constant of our Universe is extremely small ($\\approx{\\cal O}(10^{-47})\\ \\text{GeV}^4$), and assuming a reasonable $M_5$ value of ${\\cal O}(10^{19})$ GeV, the first two boundary conditions in relation (\\ref{junctions}) provide in general a huge value for $e^{B(0)}$ ($\\approx{\\cal O}(10^{45})$). This is in consistency with the scaling transformation of \\cite{brcod}, which allows us, in a solution, to scale the parameters by $e^{-S}$ and add to the warp factor the constant $S$, and acquire another solution. Therefore, the extremely small cosmological constant of the observable universe leads the cyclicity period to be around $T\\approx{\\cal O}(10^{13})$ years and the maximum scale factor value, given by (\\ref{minmax}), to be $a_{max}\\approx {\\cal O}(10^{28})$ m (where the decimal exponents in these rough estimations can vary by 1 or 2, depending on $B'(0)$ and $\\theta$ values). Luckily enough, the smallness of the cosmological constant excludes oscillatory models with small periods in astronomical terms. In more foundational words, the reason that made the cosmological constant that small, is the same that makes the cycle period and the size of the Universe that large. In this work we have been restricted to stationary solutions, where the subclass of them that possesses $H^2<0$ corresponds to eternal cyclic behavior with constant period. Numerical investigation of the full dynamics seem to consist of such stationary solutions and the transitions between them \\cite{brcod,Tetradis01}. In such transitions $H_0^2$ on the physical brane can chance sign, leading to a form of ``chaotic cyclicity\", where large intervals of (non-periodic in general) oscillatory behavior could be followed by large intervals of conventional evolution and vice versa. In this case, an initial Big Bang and/or a final Big Rip or Big Crunch (in conventional terms) could be possible. Another interesting possibility would be the exploration of our model with cosmological constants being piecewise constant functions of time, reflecting cosmological phase transitions, which could also lead to chaotic cyclicity. Note however that numerical confirmation of such behaviors is very hard due to the small probability of cyclic stationary solutions ($\\approx 10^{-2}\\%$ as we have already mentioned). These subjects are under investigation. In order for a model to serve as a description of nature, it has to explain the basic physical key issues. Especially for cyclic cosmology, amongst others these are the entropy evolution and, probably the most pressing issue, that of a fuller understanding of the bounce and the handling of the singularity. Our model provides a consistent background for cyclicity and it reveals how such a behavior arises from the full $5D$ dynamics. However, since braneworlds and brane cosmology in general arise as limits of a multi-dimensional theory unknown up to now, the $5D$ results have a phenomenological character and must be considered from this point of view. Definitely, a complete explanation and apprehension, and a successful confrontation of the aforementioned subjects, can only come through a higher-dimensional, fundamental theory of nature. For the moment we have to rely on the relevant research on cyclic cosmology, linearized gravity, M-theory and strings, which has improved our knowledge on these issues. These results can be embodied in our analysis. The most hopeful effort is the use of quantum fluctuations in order to tame the singularity, which effectively is translated into a modification of gravity by the scalar field \\cite{Bojowald01,ekpyrotic3,ekpyrotic4}. Alternatively, using loop quantum gravity we could modify non-perturbatively the dynamical equations leading to a singularity resolution as in \\cite{Maartens}. Concerning the entropy, we could include the relevant discussion in our investigation. The argument of the authors of \\cite{Turok.sing,Turok.simplified} about maximum amount of entropy possible in de Sitter spacetime, may lead our model to have a maximum cycle number between $10^{20}$ and $10^{30}$. However, the idea of the causal patch \\cite{Turok.sing} is probably the best way of handling the entropy problem so forth, and there are some interesting recent works on the subject which give a boost on cyclic cosmology \\cite{entropy}. Let us close this discussion section with some comments on the role of the brane tensions and of the bulk cosmological constant in our model. As can be numerically confirmed, setting them to zero makes it almost impossible to satisfy the boundary conditions obtaining $H^2<0$ and singularity absence in [0,1] (this can be achieved only through a careful fine-tuning since our random choice procedure gives an one-digit number of such solutions in $10^6$ parameter multiplets). On the other hand, as we showed in \\ref{B}, in the case where $\\Lambda$, $\\lambda_0$ and $\\lambda_1$ are the only non-zero parameters, an $\\approx10^{-1}\\%$ of the random parameter choices, or $\\approx6\\%$ of the solutions, correspond to $H^2<0$. In mathematical terms, $\\Lambda$, $\\lambda_0$ and $\\lambda_1$ are requisite in order to acquire a solution with $H^2<0$ in the full dynamics, in a natural and not in a fine-tuning way. In terms of physics, it is the dark energy that lies in the background of the oscillatory mechanism and allows for cyclicity to realize. Adding the fact that it determines the cycles period and the maximum scale factor value, we conclude that dark energy is crucial in the described model. This brings it closer to the ekpyrotic paradigm of the literature. In this work we examine general braneworld models and we show that cyclic behavior can naturally arise from the full $5D$ dynamics. One important feature is that brane collisions are not required, on the contrary the branes remain stable, and the cyclicity takes place on the $4D$ geometry not on the extra dimension. Another significant result is that the smallness of the cosmological constant of the observable universe pushes the cyclic period and the scale factor to astronomical large values, an essential requirement for the establishment of cyclic cosmology as a realistic alternative paradigm. Furthermore, we indicate the possibility of a ``chaotic cyclicity\", that is extremely large, non-periodic, cyclic intervals followed by extremely large intervals of conventional evolution and vice versa. After these, the model shares both the advantages and disadvantages of cyclic cosmology.\\\\ \\paragraph*{{\\bf{Acknowledgements:}}} The author acknowledges partial financial support through the research program ``Pythagoras'' of the EPEAEK II (European Union and the Greek Ministry of Education)." }, "0710/0710.5096_arXiv.txt": { "abstract": "We investigate various galaxy occupation statistics of dark matter halos using a large galaxy group catalogue constructed from the Sloan Digital Sky Survey Data Release 4 (SDSS DR4) with an adaptive halo-based group finder. The conditional luminosity function (CLF), which describes the luminosity distribution of galaxies in halos of a given mass, is measured separately for all, red and blue galaxies, as well as in terms of central and satellite galaxies. The CLFs for central and satellite galaxies can be well modelled with a log-normal distribution and a modified Schechter form, respectively. About 85\\% of the central galaxies and about 80\\% of the satellite galaxies in halos with masses $M_h\\ga 10^{14}\\msunh$ are red galaxies. These numbers decrease to 50\\% and 40\\%, respectively, in halos with $M_h \\sim 10^{12}\\msunh$. For halos of a given mass, the distribution of the luminosities of central galaxies, $L_c$, has a dispersion of about $0.15$ dex. The mean luminosity (stellar mass) of the central galaxies scales with halo mass as $L_c \\propto M_h^{0.17}$ ($M_{*,c} \\propto M_h^{0.22}$) for halos with masses $M\\gg 10^{12.5}\\msunh$, and both relations are significantly steeper for less massive halos. We also measure the luminosity (stellar mass) gap between the first and second brightest (most massive) member galaxies, $\\log L_1 - \\log L_2$ ($\\log M_{*,1}-\\log M_{*,2}$). These gap statistics, especially in halos with $M_h \\la 10^{14.0} \\msunh$, indicate that the luminosities of central galaxies are clearly distinct from those of their satellites. The fraction of fossil groups, defined as those groups with $\\log L_1 - \\log L_2\\ge 0.8$, ranges from $\\sim 2.5\\%$ for groups with $M_h\\sim 10^{14}\\msunh$ to 18-60\\% for groups with $M_h\\sim 10^{13}\\msunh$. The number distribution of satellite galaxies in groups of a given mass follows a Poisson distribution, in agreement with the occupation statistics of dark matter sub-halos. This provides strong support for the standard lore that satellite galaxies reside in sub-halos. Finally, we measure the fraction of satellites, which changes from $\\sim 5.0\\%$ for galaxies with $\\rmag\\sim -22.0$ to $\\sim40\\%$ for galaxies with $\\rmag\\sim -17.0$. ", "introduction": "In recent years, the halo occupation distribution and conditional luminosity function have become powerful statistical measures to probe the link between galaxies and their hosting dark matter halos. Although these statistical measures themselves do not give physical explanations of how galaxies form and evolve, they provide important constraints on various physical processes that govern the formation and evolution of galaxies, such as gravitational instability, gas cooling, star formation, merging, tidal stripping and heating, and a variety of feedback processes. In particular, they constrain how their efficiencies scale with halo mass. The halo occupation distribution (hereafter HOD), $P(N \\vert M)$, which gives the probability of finding $N$ galaxies (with some specified properties) in a halo of mass $M$, has been extensively used to study the galaxy distribution in dark matter halos and galaxy clustering on large scales (e.g. Jing, Mo \\& B\\\"orner 1998; Peacock \\& Smith 2000; Seljak 2000; Scoccimarro \\etal 2001; Jing, B\\\"orner \\& Suto 2002; Berlind \\& Weinberg 2002; Bullock, Wechsler \\& Somerville 2002; Scranton 2002; Kang \\etal 2002; Marinoni \\& Hudson 2002; Zheng \\etal 2002; Magliocchetti \\& Porciani 2003; Berlind \\etal 2003; Zehavi \\etal 2004, 2005; Zheng \\etal 2005; Tinker \\etal 2005). The conditional luminosity function (CLF), $\\Phi(L \\vert M) {\\rm d}L$, which refines the HOD statistic by considering the average number of galaxies with luminosity $L \\pm {\\rm d}L/2$ that reside in a halo of mass $M$, has also been extensively investigated (Yang, Mo \\& van den Bosch 2003; van den Bosch, Yang \\& Mo 2003; Vale \\& Ostriker 2004, 2006; Cooray 2006; van den Bosch et al. 2007a) and has been applied to various redshift surveys, such as the 2dFGRS, the Sloan Digital Sky Survey (SDSS) and DEEP2 (e.g. Yan, Madgwick \\& White 2003; Yang \\etal 2004; Mo et al. 2004; Wang \\etal 2004; Zehavi \\etal 2005; Yan, White \\& Coil 2004). These investigations demonstrate that the halo occupation statistics are very useful in establishing and describing the connection between galaxies and dark matter halos. Furthermore, they also indicate that the galaxy/dark halo connection can provide important constraints on cosmology (e.g.,van den Bosch, Mo \\& Yang 2003; Zheng \\& Weinberg 2007). Finally, the HOD/CLF framework also allows one to split the galaxy population in centrals and satellites, and to describe their properties separately (e.g. Cooray 2005; White \\etal 2007; Zheng \\etal 2007). As has been pointed out in Yang et al. (2005c; hereafter Y05c), a shortcoming of the HOD/CLF models is that the results are not completely model independent. Typically, assumptions have to be made regarding the functional form of either $P(N \\vert M)$ or $\\Phi(L \\vert M)$. Moreover, in all HOD/CLF studies to date, the occupation distributions have been determined in an indirect way: the free parameters of the assumed functional form are constrained using {\\it statistical} data on the abundance and clustering properties of the galaxy population. One may hope to circumvent this shortcoming by directly measure the dark matter distribution around galaxies. Such measurements can in principle be obtained through gravitational lensing and X-ray observations. However, both methods are hampered by requirements on the data quality and uncertainties related to the interpretation of the data. For instance, weak lensing measurements, which requires high-quality imaging, typically needs to resort to the stacking of many lens galaxies in order to get a detectable signal, but this stacking severely complicates the interpretation in terms of the halo masses of the lens galaxies. In the case of X-ray observations, robust constraints can only be obtained for massive clusters, but even here the interpretation of the data can be complicated due to the presence of substructure and deviations from hydrostatic equilibrium. An alternative method to directly probe the galaxy - dark halo connection is to use galaxy groups as a representation of dark matter halos and to study how the galaxy population changes with the properties of the groups (e.g., Y05c; Zandivarez et al. 2006; Robotham et al. 2006; Hansen \\etal 2007). Recently, we have constructed a large galaxy group catalogue based on the Sloan Digital Sky Survey Data Release 4 (SDSS DR4), using an adaptive halo-based group finder (Yang \\etal 2007; Paper I; Y07 hereafter). Detailed tests with mock galaxy catalogues have shown that this group finder is very successful in associating galaxies according to their common dark matter halos. In particular, the group finder performs reliably not only for rich systems, but also for poor systems, including isolated central galaxies in low mass halos. This makes it possible to study the galaxy-halo connection for systems covering a large dynamic range in masses. Various observational selection effects have been taken into account, especially the survey edge effects and fiber collisions. The halo masses for the groups are estimated according to the abundance match, using the characteristic group luminosity and stellar masses (see \\S\\ref{sec_data} below). According to tests with mock galaxy catalogues, the halo masses are estimated with a standard deviation of about 0.3 dex. With these well-defined galaxy group catalogues, one can not only study the properties of galaxies in different groups (e.g. Y05c; Yang \\etal 2005d; Collister \\& Lahav 2005; van den Bosch \\etal 2005; Robotham \\etal 2006; Zandivarez \\etal 2006; Weinmann \\etal 2006a,b; van den Bosch \\etal 2007b; McIntosh \\etal 2007), but also probe how dark matter halos trace the large-scale structure of the Universe (e.g. Yang \\etal 2005b, 2006; Coil \\etal 2006; Berlind \\etal 2007; Wang et al. 2007 in preparation). In this paper, which is the second in a series, we use the SDSS DR4 group catalogue to probe various occupation statistics and measure the CLFs for different populations of galaxies. This paper is organized as follows: In Section~\\ref{sec_data} we describe the data (galaxy and group catalogues) used in this paper. Section~\\ref{sec_CLFs} presents our measurement of the CLFs for all, red and blue galaxies. Sections~\\ref{sec_central}, ~\\ref{sec_HOD} and ~\\ref{sec_satfrac} describe the properties of central galaxies, the halo occupation statistics and the fraction of satellite galaxies, respectively. Finally, we summarize our results in Section~\\ref{sec_summary}. Throughout this paper, we use a $\\Lambda$CDM `concordance' cosmology whose parameters are consistent with the three-year data release of the WMAP mission: $\\Omega_m = 0.238$, $\\Omega_{\\Lambda}=0.762$, $n_s=0.951$, $h=0.73$ and $\\sigma_8=0.75$ (Spergel et al. 2007). \\begin{figure} \\plotone{f1.eps} \\caption{The color-magnitude relation for galaxies in our group sample. The open circles indicate the Gaussian peaks of the bi-normal distribution of galaxies in each luminosity bin. The solid dots indicate the corresponding averages of the two Gaussian peaks. The solid line is the best-fit quadratic relation to these averages (see eq.~[\\ref{quadfit}]), which we use to split the galaxies into red and blue population (color-coded accordingly).} \\label{fig:data_color} \\end{figure} \\begin{figure*} \\plotone{f2.eps} \\caption{The conditional luminosity functions (CLFs) of galaxies in groups of different mass bins. Symbols correspond to the CLFs obtained using $M_L$ as halo mass (estimated according to the ranking of the characteristic group luminosity), with solid and open circles indicating the contributions from central and satellite galaxies, respectively. The errorbars reflect the 1-$\\sigma$ scatter obtained from 200 bootstrap samples. The solid lines indicate the related best-fit parameterizations using equation~[\\ref{eq:CLF_fit}]. For comparison, we also show, with dashed lines, the CLFs obtained using $M_S$ as halo mass (estimated according to the ranking of the group's characteristic stellar mass). Results shown in this plot are obtained from Sample II. } \\label{fig:CLF} \\end{figure*} \\begin{figure*} \\plotone{f3.eps} \\caption{Similar to Fig.~\\ref{fig:CLF}, but here we show the CLFs for red (dashed lines) and blue (dotted lines) galaxies, for groups with halo masses $M_L$. The solid lines indicate the best-fit parameterizations for the CLFs of red galaxies. In both cases the central and satellite components of the CLFs are indicated separately. For clarity, the errorbars, again obtained using 200 bootstrap samples, are only shown for the red galaxies. } \\label{fig:CLF_color} \\end{figure*} \\begin{figure*} \\plotone{f4.eps} \\caption{The best fit parameters ($\\phi_s^{\\star}$ upper row, $\\alpha_s^{\\star}$ second row, $L_c$ third row, and $\\sigma_c$ bottom row) to the CLFs shown in Figs.~\\ref{fig:CLF} and ~\\ref{fig:CLF_color}, as functions of halo mass. Panels on the left, in the middle, and on the right show results for all, red, and blue galaxies, respectively. Since we have two different halo mass estimators ($M_L$ and $M_S$) and two main group samples (II and III), we have obtained CLFs for four different combinations of sample and group mass estimator. The results for all four combinations are shown using different symbols and line-styles, as indicated. The errorbars in the first two and last rows indicate the 1-$\\sigma$ variances obtained from our 200 bootstrap samples. In the third row of panels, however, the errorbars correspond to the log-normal scatter, $\\sigma_c$, shown in the bottom row of panels. For clarity the errorbars are only shown for the `$M_L$-Sample II' case, but they are very similar for the other four cases. } \\label{fig:fit_CLF} \\end{figure*} ", "conclusions": "\\label{sec_summary} Using a large galaxy group catalogue constructed from the SDSS Data Release 4 (DR4) by Y07, we have investigated various halo occupation statistics of galaxies. In particular, we have split the galaxy population in red and blue galaxies, and in centrals and satellites, and determined the conditional luminosity functions of these varies subsamples. We have also presented luminosity gap statistics, satellite fractions, and halo occupation numbers for the galaxies in our group sample. The main results are summarized as follows: \\begin{enumerate} \\item The conditional luminosity functions for central and satellite galaxies can be well modelled with a log-normal distribution and a modified Schechter form, respectively. The corresponding best fitting parameters are listed in Table 1. \\item The average scatter of the log-normal luminosity distribution of central galaxies decreases from $\\sim 0.15$ dex at the massive end ($\\log [M_h/\\msunh] \\ga 13.5$) to $\\sim 0.1$ dex at the low mass end ($\\log [M_h/\\msunh] \\sim 12.0$). However, due to the method used to assign halo masses to the groups, at the low mass end this should be considered a lower limit on the true amount of scatter. \\item The slope of the relation between the average luminosity of central galaxies (in the $^{0.1}r$-band) and halo mass, ${\\rm d}\\log L_c/{\\rm d}\\log M_h$, decreases from $\\sim 0.68$ for $\\log [M_h/\\msunh] \\ll 12.5$ to $\\sim 0.17$ for$\\log [M_h/\\msunh] \\gg 12.5$. For the stellar masses of the central galaxies we obtain that ${\\rm d}\\log M_{*,c}/{\\rm d}\\log M_h$, decreases from $\\sim 1.83$ for $\\log [M_h/\\msunh] \\ll 12.1$ to $\\sim 0.22$ for$\\log [M_h/\\msunh] \\gg 12.1$. \\item The halo (group) occupation numbers of satellite galaxies accurately follow Poisson statistics. Since the same applies to dark matter sub-halos, this supports the standard picture that satellite galaxies are associated with dark matter sub-halos. \\item In massive halos with masses $M_h\\ga 10^{14}\\msunh$ roughly 85\\% (80\\%) of the central (satellite) galaxies are red. These red fractions decrease to 50\\% (40\\%) in halos with masses $M_h \\sim 10^{12}\\msunh$. \\item By comparing the scatter in the luminosities of BCGs to the luminosity difference between the BCG and its brightest satellite, we find that the BCGs form a `special' subclass, in that their luminosities can not be considered the extreme values of the distribution of satellite luminosities, expecially in halos with masses $M_h \\la 10^{14.0} \\msunh$. \\item The fractions of fossil groups, which are defined as groups a with luminosity gap $\\log L_1 - \\log L_2 \\ge 0.8$, decreases with increasing of halo mass from 18\\%-60\\% in halos with $M_h \\sim 10^{13}\\msunh$ to $\\sim 2.5\\%$ in halos with $M_h \\sim 10^{14}\\msunh$. \\item The satellite fractions obtained from our group catalogue as functions of both luminosity and stellar mass (listed in Table~\\ref{tab:fsat}) are in good agreement with independent constraints from analyses of galaxy clustering and galaxy-galaxy lensing. \\end{enumerate} These results can be used to constrain the various physical processes related to galaxy formation and to interpret the various statistics used to describe large scale structures (e.g., galaxy correlation functions, pairwise velocity dispersions, etc.). Most of our findings are in good agreement with previous studies (e.g. Y05c, Zandivarez et al. 2006; Robotham et al. 2006) and can be linked to the semi-analytical modelling of galaxy formations (e.g., Kang et al. 2005; Zheng et al. 2005; Bower et al. 2006; Croton et al. 2006; De Lucia et al. 2007). The luminosity and stellar mass gap can be used to probe the specific formation properties of central galaxies (e.g., Vale \\& Ostriker 2007). The fraction of the red and blue populations for central and satellite galaxies can be used to probe the color evolution of satellite galaxies (ven den Bosch et al. 2007b)." }, "0710/0710.3165_arXiv.txt": { "abstract": "Stellar-mass black holes are discovered in X-ray emitting binary systems, where their mass can be determined from the dynamics of their companion stars\\cite{rem06,cha06,oro03}. Models of stellar evolution have difficulty producing black holes in close binaries with masses \\boldmath{$>10\\,M_{\\odot}$} (ref.\\ 4), which is consistent with the fact that the most massive stellar black holes known so far\\cite{cha06,oro03} all have masses within \\boldmath{$1\\sigma$} of \\boldmath{$10\\, M_{\\odot}$}. Here we report a mass of \\boldmath{$15.65 \\pm 1.45\\,M_{\\odot}$} for the black hole in the recently discovered system M33 X-7, which is located in the nearby galaxy Messier 33 (M33) and is the only known black hole that is in an eclipsing binary\\cite{pie06}. In order to produce such a massive black hole, the progenitor star must have retained much of its outer envelope until after helium fusion in the core was completed\\cite{bro01}. On the other hand, in order for the black hole to be in its present 3.45 day orbit about its \\boldmath{$70.0 \\pm 6.9 M_{\\odot}$} companion, there must have been a ``common envelope'' phase of evolution in which a significant amount of mass was lost from the system\\cite{tau06}. We find the common envelope phase could not have occured in M33 X-7 unless the amount of mass lost from the progenitor during its evolution was an order of magnitude less than what is usually assumed in evolutionary models of massive stars\\cite{sch92,mey94,vaz07}. ", "introduction": " ", "conclusions": "" }, "0710/0710.1998_arXiv.txt": { "abstract": "The concept of black hole entropy is one of the most important enigmas of theoretical physics. It relates thermodynamics to gravity and allows substantial hints toward a quantum theory of gravitation. Although Bekenstein conjecture --assuming the black hole entropy to be a measure of the number of precollapse configurations-- has proved to be extremely fruitful, a direct and conclusive proof is still missing. This article computes accurately the entropy evaporated by black holes in $(4+n)$ dimensions taking into account the exact greybody factors. This is a key process to constrain and understand the entropy of black holes as the final state is unambiguously defined. Those results allow to generalize Zurek's important argument, in favor of the Bekenstein conjecture, to multi-dimensional scenarios. ", "introduction": " ", "conclusions": "" }, "0710/0710.3023_arXiv.txt": { "abstract": "Three-dimensional large eddy simulations of solar surface convection using realistic model physics are conducted. The thermal structure of convective motions into the upper radiative layers of the photosphere, the range of convection cell sizes, and the penetration depths of convection are investigated. A portion of the solar photosphere and the upper layers of the convection zone, a region extending $60\\times60$ Mm horizontally from 0 Mm down to 20 Mm below the visible surface, is considered. We start from a realistic initial model of the Sun with an equation of state and opacities of stellar matter. The equations of fully compressible radiation hydrodynamics with dynamical viscosity and gravity are solved. We use: 1) a high order conservative TVD scheme for the hydrodynamics, 2) the diffusion approximation for the radiative transfer, 3) dynamical viscosity from subgrid scale modeling. The simulations are conducted on a uniform horizontal grid of $600\\times600$, with 168 nonuniformly spaced vertical grid points, on 144 processors with distributed memory multiprocessors on supercomputer MVS-15000BM in the Computational Centre of the Russian Academy of Sciences. ", "introduction": "Convection near the solar surface has a strongly non-local and dynamical character. Hence, numerical simulations provide useful information on the spatial structures resulting from convection and help in constructing consistent models of the physical processes underlying the observed solar phenomena. We conduct an investigation of the temporal evolution and growth of convective modes on scales of mesogranulation and supergranulation in a three-dimensional computational box. In previous work by the author [\\citet{ustyugs06}] it was shown that collective motion of small convective cells of granulation expels weak magnetic field on the edges of cells at mesogranular scales. The average size of such cells is 15-20 Mm and the lifetime of order 8-10 solar hours. Simulation of solar photosphere convection [\\citet{stein06}] in a computational domain of size 48 Mm in the horizontal plane and 20 Mm in depth showed that the sizes of convective cells increase with depth. The purpose of this work is to investigate the development and scales of convection in a region of size 60 Mm in the horizontal plane and 20 Mm in depth. ", "conclusions": "" }, "0710/0710.3037_arXiv.txt": { "abstract": "I present a new census of the stellar population in the Chamaeleon~I star-forming region. Using optical and near-IR photometry and followup spectroscopy, I have discovered 50 new members of Chamaeleon~I, expanding the census of known members to 226 objects. Fourteen of these new members have spectral types later than M6, which doubles the number of known members that are likely to be substellar. I have estimated extinctions, luminosities, and effective temperatures for the known members, used these data to construct an H-R diagram for the cluster, and inferred individual masses and ages with the theoretical evolutionary models of Baraffe and Chabrier. The distribution of isochronal ages indicates that star formation began 3-4 and 5-6~Myr ago in the southern and northern subclusters, respectively, and has continued to the present time at a declining rate. The IMF in Chamaeleon~I reaches a maximum at a mass of 0.1-0.15~$M_\\odot$, and thus closely resembles the IMFs in IC~348 and the Orion Nebula Cluster. In logarithmic units where the Salpeter slope is 1.35, the IMF is roughly flat in the substellar regime and shows no indication of reaching a minimum down to a completeness limit of 0.01~$M_\\odot$. The low-mass stars are more widely distributed than members at other masses in the northern subcluster, but this is not the case in the southern subcluster. Meanwhile, the brown dwarfs have the same spatial distribution as the stars out to a radius of $3\\arcdeg$ (8.5~pc) from the center of Chamaeleon~I. ", "introduction": "\\label{sec:intro} The characteristics of the distributions of masses, ages, and positions in a newborn stellar population are determined by the process of star formation. As a result, measurements of these distributions in star-forming regions are potentially valuable for testing models of the birth of stars and brown dwarfs. For instance, the properties of the stellar initial mass function \\citep[IMF,][]{mey00} can constrain the relative importance of turbulent fragmentation \\citep{pn02}, gravitational fragmentation \\citep{lar85}, dynamical interactions \\citep{bon03}, and accretion and outflows \\citep{af96} in regulating the final masses of stars. Models of the star formation rates of molecular clouds (e.g., constant, accelerating, bursts) can be tested against the distributions of ages and positions of members of young clusters \\citep{fei96,pal97,har01}. The spatial distributions also provide insight into cloud fragmentation, binary formation, cluster dynamics, and the origin of brown dwarfs \\citep{lar95,hil98,har02,luh06tau}. To obtain measurements of this kind, one must identify the members of star-forming regions and estimate their masses and ages. Only a few young stellar populations have been characterized in detail, such as the Orion Nebula Cluster \\citep{hil97}, Taurus \\citep{kh95}, and IC~348 in Perseus \\citep{luh03ic}. The Chamaeleon~I star-forming region is amenable to a thorough census of its stellar population for several reasons. It is among the nearest star-forming regions \\citep[$d=160$-170~pc,][]{whi97,wic98,ber99}, exhibits less extinction than many young clusters ($A_V\\lesssim5$), is compact enough that a large fraction of the region can be surveyed to great depth in a reasonable amount of time, and is sufficiently rich for a statistically significant analysis of its stellar population. In addition, because Chamaeleon~I is relatively isolated from other star-forming regions, confusion with other young populations is minimal. Previous surveys have already identified more than 150 young stars and brown dwarfs in Chamaeleon~I through photometric variability, H$\\alpha$ emission, X-ray emission, mid-infrared (IR) excess emission, and optical and near-IR color-magnitude diagrams \\citep[][references therein]{com04,luh04cha,luh07cha}. However, in the census of known members produced by these surveys, the completeness as a function of mass and position is unknown \\citep{luh04cha}. In this paper, I present a set of magnitude-limited surveys for members of Chamaeleon~I that have well-defined completeness limits (\\S~\\ref{sec:new}). I then use the new census of Chamaeleon~I to measure the star formation history (\\S~\\ref{sec:hr}), IMF (\\S~\\ref{sec:imf}), and spatial distribution of its stellar population (\\S~\\ref{sec:spatial}). ", "conclusions": "I have presented an extensive search for new members of the Chamaeleon~I star-forming region. Because the completeness limits of my survey are well-determined, I have been able to perform robust measurements of the distributions of members of Chamaeleon~I as a function of mass, position, and age. The primary results of this study are summarized as follows: \\begin{enumerate} \\item I have discovered 50 new members of Chamaeleon~I, which increases the census of known members to 226 objects. The new members include 14 objects that are later than M6 ($M\\lesssim0.08$~$M_\\odot$) and the two faintest known members of the cluster, which may have masses of only 0.005-0.01~$M_\\odot$. The current census now contains 28 members that are likely to be substellar. \\item The distribution of isochronal ages for members of Chamaeleon~I between 0.1-1~$M_\\odot$ suggests that star formation has occurred for the past 3-4 and 5-6~Myr in the southern and northern subclusters, respectively, at rates that have declined with time. \\item The IMF in Chamaeleon~I reaches a maximum at a mass of 0.1-0.15~$M_\\odot$, which is similar to the turnover mass observed in IC~348 and the Orion Nebula Cluster \\citep{hil97,hc00,mue02,mue03,luh03ic}. The substellar IMF is roughly flat in logarithmic units and shows no indication of reaching a minimum down to a completeness limit of 0.01~$M_\\odot$. \\item Chamaeleon~I does not contain a widely-distributed population of brown dwarfs, which is contrary to the predictions of some embryo ejection models. Instead, the substellar members share the same spatial distribution as the stars. However, low-mass stars in the northern subcluster do appear to have a wider distribution than members at other masses, which resembles the mass segregation that has been previously observed in Orion and IC~348 \\citep{hc00,mue03}. \\end{enumerate}" }, "0710/0710.1518_arXiv.txt": { "abstract": "% After presenting three ways of defining a bulge component in disc galaxies, we introduce the various types of bulges, namely the classical bulges, the boxy/peanut bulges and the disc-like bulges. We then discuss three specific topics linked to bulge formation and evolution, namely the coupled time evolution of the bar, buckling and peanut strengths; the effect of velocity anisotropy on peanut formation; and bulge formation via bar destruction. ", "introduction": "Three ways of defining a bulge have been used so far, one morphological, the second photometrical and the third kinematical. Based on morphology, a bulge is a component of a disc galaxy that has a smooth light distribution that swells out of the central part of a disc viewed edge-on. This definition has the disadvantage of being applicable only to edge-on systems and the advantage of necessitating only an image of the galaxy. The second definition is based on photometry and defines a bulge as the extra light in the central part of the galaxy, above the exponential profile fitting the remaining (non central) part of the disc. In earlier papers this component was fitted with an $r^{1/4}$ law, while more recent ones use its generalisation to an $r^{1/n}$ law (S\\'ersic 1968). This definition has the advantage of being applicable to disc galaxies independent of their inclination. It has also the advantage of leading to quantitative results about the light distribution, but has the disadvantage of assigning to the bulge any extra central luminosity of the disc, independent of its origin. The third definition is based on kinematics, and in particular on the value of $V/\\sigma$, or more specifically on the location of the object on the ($V/\\sigma$, ellipticity) plot, which is often referred to as the Binney diagram (Binney 1978, 2005). This definition, potentially quite powerful, has unfortunately been very little used so far, due to the small number of galaxies for which the necessary data are available, a situation which is rapidly improving with large surveys, such as SAURON (Bacon \\tal 2001; de Zeeuw \\tal 2002; Peletier this volume). The lack of a single, clear-cut definition of a bulge is due to the fact that disc galaxies are viewed in different orientations and also to the fact that not all types of data are available for all objects. Nevertheless, it has led to considerable confusion and to the fact that bulges are an inhomogeneous class of objects. Indeed, many different types of objects, with very different properties and formation histories are included in the general term `bulges'. To remedy this, Kormendy (1993) and Kormendy \\& Kennicutt (2004) distinguish classical bulges from pseudo-bulges, the latter category encompassing all bulges that are not classical. Athanassoula (2005a) argues that pseudo-bulges also are an inhomogeneous class of objects, and thus distinguishes three types of bulges. {\\bf Classical bulges} are formed by gravitational collapse or hierarchical merging of smaller objects and corresponding dissipative gas processes. The material forming this bulge could be externally accreted, or could come from clumps in the proto-disc. In general, bulges of this type are formed before the actual disc (e.g. Steinmetz \\& M\\\"uller 1995; Noguchi 1998; Immeli \\tal 2004). Nevertheless, a bulge can also form from material externally accreted at much later stages (e.g. Pfenniger 1991; Athanassoula 1999; Aguerri, Balcells \\& Peletier 2001; Fu, Huang, Deng 2003). Their morphological, photometrical and kinematical properties are similar to those of ellipticals. {\\bf Box/peanut bulges} (B/P) form from a vertical instability of the disc material. This has often been observed in $N$-body simulations of bar-unstable discs, where the initial stage of bar formation is followed by a puffing up of the inner parts of the bar (e.g. Combes \\& Sanders 1981; Combes \\tal 1990; Raha \\tal 1991; Athanassoula \\& Misiriotis 2002; Athanassoula 2003, 2005a; O'Neil \\& Dubinski 2003; Martinez-Valpuesta \\& Shlosman 2004; Debattista \\tal 2004, 2006; Martinez-Valpuesta \\tal 2006). Viewed side-on (i.e. edge-on with the line of sight along the bar minor axis), this structure protrudes from the disc and has a characteristic boxy or peanut shape whose size is of the order of a few disc scale-lengths. Thus, a box/peanut bulge is just {\\it part} of a bar seen side-on. Finally {\\bf disc-like bulges} form from inflow of (mainly) gas material to the centre of the galaxy and subsequent star formation (e.g. Athanassoula 1992; Friedli \\& Benz 1993; Heller \\& Shlosman 1994; Wada \\& Habe 1995) . The torque exerted by the bar pushes gas, and to a lesser extent also stars, to the inner parts of the disc where they form an inner disc. Star formation can be very high there, due to the increased gas density. Thus the result of this process should be a central disc, or disc-like object, whose stellar component should include a sizable fraction of young stars and whose size should be less than, or of the order of a kpc. Such a component could harbour sub-structures such as spirals, or bars, as is indeed sometimes observed (Kormendy \\& Kennicutt 2004 and references therein). It is thus clear that disc-like bulges are very different objects from boxy/peanut bulges, since they are much smaller, have a different shape, different kinematics and provide a different type of excess on the radial photometric profiles. They also have different formation histories. The different formation histories of these three types of objects lead to different properties -- morphological, photometrical and kinematical -- which in turn help classify observed bulges into one of the three above mentioned types. Nevertheless, as stressed by Athanassoula (2005a), different types of bulges often co-exist and it is possible to find all three types of bulges in the same simulation, or galaxy. Realising the non-homogeneity of the objects classified as bulges and attempting to classify them is only the first step. Much more work is now necessary, particularly on two issues. The first one is the understanding of the formation and evolution of these types of objects. The second one is to predict the properties of these objects, starting from their formation scenarios. The latter is particularly important in order to bridge the gap between classification schemes based on formation histories and classification schemes based on observed properties. Here we make small contributions to both these issues, using $N$-body simulations. In Sect. 2 we present the time evolution of the bar, the buckling and the peanut strengths and their interplay. Sect. 3 discusses the velocity anisotropy and its link to the above strengths. Finally, in Sect. 4 we discuss the photometrical properties of a destroyed bar and boxy/peanut bulge. ", "conclusions": "" }, "0710/0710.3347_arXiv.txt": { "abstract": "It has been suggested that fossil groups could be the canibalized remains of compact groups, that lost energy through tidal friction. However, in the nearby universe, compact groups which are close to the merging phase and display a wealth of interacting features (such as HCG 31 and HCG 79) have very low velocity dispersions and poor neighborhoods, unlike the massive, cluster-like fossil groups studied to date. In fact, known z$=$0 compact groups are very seldom embedded in massive enough structures which may have resembled the intergalactic medium of fossil groups. In this paper we study the dynamical properties of CG6, a massive compact group at z$=$0.220 that has several properties in common with known fossil groups. We report on new \\sloang~and \\sloani~imaging and multi-slit spectroscopic performed with GMOS on Gemini South. The system has 20 members, within a radius of 1 h$_{70}^{-1}$ Mpc, a velocity dispersion of 700~\\kms~and has a mass of 1.8 $\\times$ 10$^{14}$ h$_{70}^{-1}$~\\Msol, similar to that of the most massive fossil groups known. The merging of the four central galaxies in this group would form a galaxy with magnitude $M_{r'} \\sim -23.4$, typical for first-ranked galaxies of fossil groups. Although nearby compact groups with similar properties to CG 6 are rare, we speculate that such systems occurred more frequently in the past and they may have been the precursors of fossil groups. ", "introduction": "\\label{sec:intro} Groups of galaxies are small systems of typically a few L$^{*}$ galaxies, which comprise over 55\\% of the nearby structures in the universe. A small fraction of galaxy groups are classified as {\\it compact} groups, which are responsible for $\\sim$ 1\\% of the luminosity density of the universe \\citep{mdoh91}. Although they are rare objects in the nearby universe, their high galactic densities and low velocity dispersions make them ideal systems for the study of galaxy transformation through galaxy-galaxy collisions. As expected, these systems have a high fraction of interacting members, although merged objects are rare \\citep{zepf93}. They are commonly believed to evolve through dynamical friction and finally merge to form one single galaxy \\citep{Barnes92}. \\citet{vikhlinin99} and \\citet{jones03} have suggested that the merging of compact groups can lead to the formation of fossil groups. A fossil group (FG, hereafter) is a system with an extended and luminous X-ray halo (L$_X$ $>$ $10^{42}$ h$_{70}^{-2}$ erg s$^{-1}$), dominated by one single brighter than L$^{*}$ elliptical galaxy, surrounded by low-luminosity companions \\citep[where the difference in magnitude between the bright dominant elliptical and the next brightest companion is $>$ 2 mag in the R-band;][]{jones03}. One important goal of this article is to investigate if compact groups (CG, hereafter) as we known them today, could be the precursors of FGs. In order to answer this question, we summarize, in section 2, the properties of a few of the most strongly interacting nearby CGs known, which are about to merge, and in section 3, we describe the properties of the five FGs which have been studied spectroscopically so far. In section 4, we present new observations for a CG embedded in a cluster-size potential, at redshift z$=$0.22 and section 5 puts together all the observations described and discusses the CG-FG scenario. Throughout this paper we adopt when necessary a standard cosmological model: $H_{0}=70\\,h_{70}\\,$km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{m}=0.3$ and $\\Omega_{\\Lambda}=0.7$. At z$=$0.22, 1\\arcsec~corresponds to $ 3.5\\, h_{70}^{-1}\\,$kpc. \\section {Interacting compact groups: HCG 31, HCG 79 and HCG 92} \\label{sec:hcg} There is evidence from both observations and simulations that groups evolve through dynamical friction and coalesce to form more compact structures as the universe ages. A few of the most compact, and therefore most evolved groups known, from Hickson's catalogue \\citep{h92} are HCG 31, HCG 79 (or Seyfert Sextet) and HCG 92 (or Stephan's quintet). The study of these groups is important to help understanding processes common in merging systems, environments that may have occurred more often in the high-redshift universe. HCG 31 is a group at z$\\sim$0.013 and with a velocity dispersion of $\\sigma$ $\\sim$ 60 \\kms. This is a gas-rich group with intense star forming activity \\cite[e.g. ][]{moetal06, amram07}, dominated by a central pair of interacting dwarf galaxies A$+$C. HCG 31 is thought to be in a pre-merging phase \\citep{amram05,vm05} and it has well developed tidal tails seen in H$\\alpha$ and HI. The group hosts two excellent candidates for tidal dwarf galaxies, namely member F, in the southern tail and member R, 50 h$_{70}^{-1}$ kpc to the north of the group (for an assumed distance modulus of DM=33.8). HCG 79, also known as {\\it Seyfert Sextet}, was originally identified as a sextet of galaxies but it is now known to be a quartet at z$=$ 0.0145 (the 5th object is in the background and the 6th is a luminous tidal debris to the northeast of the group). This is the most CG in Hickson's catalogue with a galaxy-galaxy distance below 10 kpc (for an adopted DM=34.0) and a velocity dispersion of $\\sigma$ = 138 \\kms. The four galaxies present morphological distortions and increased activity (tidal debris, bar in HCG 79B, dust lane in HCG 79A, radio and infrared emission, disturbed rotation curves and nuclear activity). The group presents a prominent intra-group light envelope which contains 45\\% of the total light of the group \\citep{darocha05} and irregular envelopes of HI \\citep{wil91} and X--rays \\citep{pil95}. These suggest that recent or on-going interaction is taking place within this system. HCG 92, also known as Stephan's quintet, is in reality a quartet with z$=$ 0.0215 and a foreground galaxy. It is the most well studied CG -- multi wavelength data are available from radio to X-rays. Most of the gaseous material in Stephan's quintet is concentrated not in the galaxies but in the intragroup medium, suggesting that collisions among group members may have happened frequently. A number of tidal dwarf galaxies and intergalactic HII regions have been identified in this group \\citep[e.g.][]{moetal01,moetal04,xu05}. Of the three groups described above, HCG 92 is the only one to have detected X-rays, with a total bolometric luminosity of 2.96$\\times$10$^{42}$ h$^{-2}_{70}$ \\ergs~\\citep{Xue2000}. These three spiral-rich groups are thought to be in their final stages of evolution -- they are, in fact, some of the most compact systems found in the Hickson's catalogue. Yet, they have members that can be clearly identified as individual galaxies, suggesting that once merging starts, it may proceed quickly, and the groups may no longer be recognized as such. The bright members of these groups will almost certainly end up as a single galaxy pile. A discussion of whether these systems will most likely end up as FGs or as single isolated elliptical galaxies is deferred to section 5. In the following section, some of the optical properties of the FGs studied so far are summarized. ", "conclusions": "Dynamical friction and subsequent merging are probably the processes responsible for the lack of L$^{*}$ galaxies in FGs. Considering the merging scenario, it is possible that the overluminous central galaxy in a FG has been formed within a substructure, inside a larger structure. In that case, one could think of a scenario where a CG was formed within a rich group, which would then have merged, leaving behind the brightest elliptical galaxy of what today is seen as a FG. One weak argument against this scenario is that the nearby examples of CGs are not usually found within such massive structures, but instead are more often surrounded by very sparse structures. There are, however, examples such as CG6, surrounded by large numbers of lower-luminosity galaxies, inhabiting a deep potential well. We would like to test the hypothesis that CGs, as observed in the nearby universe, could be the precursors of FGs. We may examine two aspects: (1) if the sum of the luminosities of the brightest CG galaxies is similar to the luminosity of a first-ranked FG galaxy and; (2) if the neighbourhoods of CGs are rich, i.e., if the system as a whole (group plus environment) has a velocity dispersion/mass similar to that of a FG. We compute the total luminosity of the galaxies in the soon-to-merge CGs, HCG 31, HCG 79 and HCG 92, to check how these compare with the luminosities of first-ranked galaxies in FGs. Adding up the luminosities of galaxies HCG 31 A to C, G and Q, which are the brightest in the group HCG 31, a magnitude of M$_R = -22.5$ is obtained (for a distance modulus, DM = 33.8). Summing up the luminosities of galaxies HCG 79 A-D, an equal total magnitude of M$_R=-22.5$ is obtained (for DM = 34.0). These are upper limits on the luminosities of these objects given that several members are starburst galaxies. After fading, the merged central object in HCG 31 and HCG 79 will have somewhat lower magnitudes than that of a typical first-ranked galaxy in a FG. Fossil groups first-ranked galaxies have luminosities well above L$^{*}$. For the five FGs studied by \\citet{jones03}, the first-ranked galaxies had a median luminosity of $M_R=-23.2$ and for the 34 FGs found in the SDSS DR5 by \\citet{dosSantos07} the median luminosity was $M_R=-23.5$. Although for HCG 92, the final object (adding up luminosities of galaxies A-E) would have an absolute magnitude of $M_R=-24.2$ (for DM$=$34.8), which after allowing for some fading, could be similar to that of an FG first-ranked galaxy, HCG 92 would possibly still not resemble an FG when merged, because its neighbourhood is very sparse, i.e., it is not embedded in any larger structure, as it is often the case for the central galaxy in FGs. This is in agreement with its relatively low bolometric X-ray luminosity of 2.96 $\\times$ 10$^{42}$ h$_{70}^{-2}$ ergs s$^{-1}$ \\citep{Xue2000}. The environments of nearby CGs have been surveyed by \\citet{ribeiro98,zabludoff98,carrasco06} among others. Spectroscopy of dozens of members in the neighbourhood of quite a number of groups was obtained, confirming in all cases that CGs have low velocity dispersions typical of the group regime (typically 200-300 km s$^{-1}$). In fact, even for HCG 62, thought to be one of the most massive CGs in Hickson's catalogue, the velocity dispersion obtained from 45 members of the system showed that it is a bonafide group (376 km s$^{-1}$). HCG 62 was suggested by \\citet{ponman93} as an example of a system that could turn into a FG in a few Gygayears, but its velocity dispersion is still much lower than the value of $\\sim$ 600 km s$^{-1}$, typical for rich FGs. Two other massive nearby CGs in Hickson's catalogue are HCG 94 and HCG 65. The first is known to have a very high bolometric X-ray luminosity of 2.35 $\\times$ 10$^{44}$ h$_{70}^{-2}$ \\citep{Xue2000} which may, however, be contaminated by the emission of a nearby cluster. The velocity dispersion obtained from 11 members in this system gives a value of 479 km s$^{-1}$. HCG 65 is the center of the cluster Abell 3559. It is in the heart of the Shapley supercluster and its location makes it hard to disentangle its dynamics and determine its mass. The three most massive Hickson CGs known, HCG 62, HCG 65 and HCG 94, are strongly early-type dominated, as expected from the velocity dispersion-morphology relation observed for CGs. The conclusion is then that a typical CG, as observed at z=0, is unlike to turn into a FG. It is more likely to merge into an isolated elliptical galaxy. For the compact group CG 6, at $z=0.22$ and $\\sigma =$703 km s$^{-1}$, if we merge the four central galaxies (A, B, C and D), we end up with a galaxy with total magnitude \\sloani$=$16.31 and \\sloang$=$17.80 mag (M$_{i'}$=--23.87 and M$_{g'}$=--22.38, with no k-correction). Using the color relation for a galaxy at z$=$0.2 from \\cite{fuk95}, the magnitude in r$^{'}$ will be 16.83 or M$_{r}=-23.35$. This magnitude is similar to those of typical central galaxies in FGs and the velocity dispersion of the system is typical for the studied FGs. However, no gap of at least two magnitudes between the first-ranked relic and the remaining objects of the system would be observed because there is at least one other bright galaxy in the system within half the virial radius of the group. We point out that one other example of a possible massive system, at z=0.39, which may turn into a FG, has recently been discovered by \\cite{rines07}. Spectroscopic studies of CGs at medium redshifts may find many more of such objects." }, "0710/0710.1497_arXiv.txt": { "abstract": "We present the results of an analysis of a well-selected sample of galaxies with active and inactive galactic nuclei from the Sloan Digital Sky Survey, in the range $0.01 < z < 0.16$. The SDSS galaxy catalogue was split into two classes of active galaxies, Type~2 AGN and composites, and one set of inactive, star-forming/passive galaxies. For each active galaxy, two inactive control galaxies were selected by matching redshift, absolute magnitude, inclination, and radius. The sample of inactive galaxies naturally divides into a red and a blue sequence, while the vast majority of AGN hosts occur along the red sequence. In terms of H$\\alpha$ equivalent width, the population of composite galaxies peaks in the valley between the two modes, suggesting a transition population. However, this effect is not observed in other properties such as colour-magnitude space, or colour-concentration plane. Active galaxies are seen to be generally bulge-dominated systems, but with enhanced H$\\alpha$ emission compared to inactive red-sequence galaxies. AGN and composites also occur in less dense environments than inactive red-sequence galaxies, implying that the fuelling of AGN is more restricted in high-density environments. These results are therefore inconsistent with theories in which AGN host galaxies are a `transition' population. We also introduce a systematic 3D spectroscopic imaging survey, to quantify and compare the gaseous and stellar kinematics of a well-selected, distance-limited sample of up to 20 nearby Seyfert galaxies, and 20 inactive control galaxies with well-matched optical properties. The survey aims to search for dynamical triggers of nuclear activity and address outstanding controversies in optical/IR imaging surveys. ", "introduction": "Active Galactic Nuclei (AGN) have long been considered a curiosity in their own right \\citep{schmidt63, lybell69,blan74}, but are now recognised to be integral to galaxy formation and evolution. At early cosmological epochs, gas-rich galaxies are thought to form and collide in violent mergers \\citep{mihos96}, triggering vast bursts of star formation, and fuelling supermassive black holes in their cores \\citep{kriek06}. Recent simulations of isolated and merging galaxies by \\cite{spring05} have incorporated feedback from star formation and black-hole accretion, and find that once an accreting supermassive black hole (SMBH) has grown to some critical size, the AGN feedback terminates its growth as a large fraction of the remaining nuclear gas is driven out by the powerful quasar. In the current epoch, the peak of the quasar era is over and the galaxy merger rate has declined \\citep{struck97}. Nevertheless, at least 20\\% of today's galaxies show scaled-down quasar activity in their centres, and direct measurements of active and inactive galaxy dynamics have revealed a tight correlation between the central black-hole mass and host galaxy stellar velocity dispersion or bulge mass. This $M_{\\bullet}-\\sigma$ relation \\citep{geb00,merr01} points to an intimate link between host galaxy evolution and central black-hole growth, suggesting that all bulge-dominated galaxies today harbour dead quasars and that a lack of nuclear activity cannot be attributed to the absence of a central black hole. Therefore, given the ubiquity of supermassive black holes, what determines the degree of nuclear activity in todays galaxies and what role is played by the host galaxy in triggering and fuelling their dormant black holes remain open issues. Previously, studies of nearby active galaxies were based on small galaxy samples, like the C$f$A Seyfert Sample \\citep{huchra92}, the 12 Micron Active Galaxy Sample \\citep{rush93} and the \\cite{ho95} galaxy sample of approximately 486 galaxies covering a range of activity types. Now, however, the standard has been set by the Sloan Digital Sky Survey (SDSS), from which a range of useful samples of both active and inactive galaxies can be selected in a well-defined, uniform way. The SDSS \\citep{york00} provides, for the first time, 5-band photometry and spectroscopy of many thousands of low redshift AGN \\citep{kauf03c,hao05a}, enabling constraints to be placed on galaxy and AGN evolution over a wide range of galaxy masses, and acts as the definitive supporting data to all detailed follow-up galaxy studies. Initial investigations into the growth and evolution of black holes using these databases have yielded contrasting results --- the origin of the correlation between galaxy bulge and central black-hole masses is hotly debated in particular. \\citet{heck04} find that the majority of present-day accretion occurs onto 10$^{8}$ solar-mass black holes in moderate mass galaxies, suggesting that bulge and black-hole evolution is still tightly coupled today, and that the evolution of AGN luminosity functions is driven by a decrease in the mass scale of accreting black holes. In contrast, the \\citet{hao05b} study of about 3\\,000 SDSS AGN concludes that evolution in AGN luminosity functions is driven by evolution in the Eddington ratio, rather than black-hole mass. Meanwhile, \\cite{grup04} and \\cite{grup05} use a sample of 75 X-ray selected AGN to argue that black holes in Narrow Line Seyfert 1 galaxies in particular, grow by accretion in well-formed bulges to produce the $M_{\\bullet}-\\sigma$ relation over time, refuting theories for the origin of black-hole / bulge relations in which black-hole mass is a constant fraction of bulge mass at all epochs or in which bulge growth is controlled by AGN feedback \\citep{king03,ferr02,kriek06}. The subclass of NLS1s has been further explored more recently by \\cite{komo07}, who find that NLS1s are accreting at a rate higher than the Eddington rate, confirming that their BHs must be growing. They suggest that either NLS1 galaxies evolve into Broad Line Seyfert 1 (BLS1) galaxies with respect to their black hole mass distribution, which would require some change in the bulge properties, possibly due to feedback, or that NLS1s are just low-mass extensions of BLS1 galaxies, and the high accretion rate could just be caused by a relatively short-lived accretion phase. Another characteristic revealed by large automated surveys such as the SDSS, is that the galaxy population is found to be bimodal in colour \\citep{lilly95,strat01}, with it now being more natural to describe a galaxy as being on the ``red sequence'' or ``blue sequence'', rather than being ``early type'' or ``late type'' \\citep{bald06}. A key goal of galaxy evolution theory is then to explain the colour bimodality of galaxies, the relationships within each sequence, and where active hosts fit into this picture. Therefore, the understanding of galaxy formation and evolution processes necessitates full inclusion of AGN and their hosts. Recently efforts have been made to try to understand where AGN host galaxies fall in colour-magnitude space. The red sequence consists of mainly massive, passively evolving galaxies, while the majority of galaxies show blue colours, attributed to ongoing star formation. The two sequences are separated by a relatively narrow valley in colour space. The emerging consensus is that intense star formation is fuelled by galaxy mergers at high redshift, which forms massive bulges, and at some point the star formation ceases, resulting in the galaxy migrating from the blue sequence to the red sequence. AGN have been singled out as the mechanism for star-formation quenching, leading to suggestions that AGN should occupy a distinct, `transition' region, of the colour-magnitude diagram (CMD). However this does not necessarily have to be the case, as the resulting galaxy could just be bluer due to increased star-burst activity, or redder due to enhanced dust (e.g., Luminous Infra-Red Galaxies), or some combination of the original colours of the merging galaxies. Young stellar populations and nuclear starbursts are therefore also important components in the dynamics of nuclear activty \\citep{gonza98,sarzi07}. \\cite{nand07} studied the colour-magnitude relation for a sample of 50 X-ray selected AGN from the AEGIS survey (All-wavelength Extended Groth strip International Survey), in the range $0.6 < z < 1.4$. They conclude that AGN fall on the red sequence or on the red edge of the blue sequence, with many in between these two modes. \\cite{mart07} further explored the idea of a `transition' region by exploiting the separation of the blue and red sequences in the ${\\rm UV} - r$ colour-magnitude diagram. Using a sample of UV selected galaxies from the GALEX Medium Imaging Survey, along with SDSS data, they explored the nature of galaxies in the transition zone. The AGN fraction of their sample was found to peak in the transition zone, and they also found circumstantial evidence that star-formation quenching rates were higher in higher luminosity AGN. Higher image quality than the SDSS is provided by the Millennium Galaxy Catalogue (MGC; \\citealt{liske03,cross04}), which has obtained redshifts for 10\\,095 galaxies to $B < 20 {\\rm\\,mag}$, covering 37 deg$^{2}$ of equatorial sky. The MGC study of galaxy bimodality was in the colour-structure plane, (\\citealt{drive06}; hereafter D06) and expressed a contrasting theory for the bimodality. The MGC survey has sufficient resolution to achieve reliable bulge-disc decomposition \\citep{allen06}, and in separating out the two components, D06 claim that galaxy bimodality is caused by the bulge-disc nature of galaxies, and not by two distinct galaxy classes at different evolutionary stages. In this case, they show that the bulge-dominated, early-type galaxies populate one peak and the bulge-less, late-type galaxies occupy the second. The early- and mid-type spirals sprawl across and between the peaks. They also propose that the reason for the dual structure of galaxies is due to galaxy formation proceeding in two stages; first, there is an initial collapse phase, which forms the centrally concentrated core and a black hole, and second, there is the formation of a planar rotating disc caused by accretion of external material building up the galaxy disk. In this paper, we present the properties of a magnitude-limited sample of ``active'' and ``inactive'' galaxies carefully selected from the SDSS, which also acts as the parent sample for detailed followup studies using the IMACS-IFU on the Magellan 6.5m telescope. Our sample selection is described in \\S 3, the sample properties in \\S 4, and implications of these properties in \\S 5. We conclude by introducing the Magellan survey, which is now underway. ", "conclusions": "We have carried out a robust classification of `active' and `inactive' galaxies in the SDSS, based on their emission line properties and locations in the BPT diagram. We have compared the active galaxies with well selected control galaxies matched in redshift, absolute magnitude, aspect ratio and radius. We find that: \\begin{itemize} \\item Type~2 AGN host galaxies occur mainly on the red sequence of the CMD, but show increased levels of H$\\alpha$ flux in their emission compared to inactive red-sequence galaxies (Fig.~\\ref{fig:H-alpha}(a) and Fig.~\\ref{fig:CMD}(a)); \\item A separate class of composite galaxies appears to peak on the blue edge of the red sequence on the CMD, whereas the peak of the H$\\alpha$ distribution places composite galaxies firmly in the valley between the blue and red sequences (Fig.~\\ref{fig:H-alpha}(b) and Fig.~\\ref{fig:CMD}(b)); \\item Colour-concentration relations, however, show a more complex, possibly double, morphology in the peak of the composite distribution, rather than being in a valley (Fig.~\\ref{fig:colour-conc-comp}); \\item AGN (and composites) are found in less dense environments on average then matched inactive red-sequence galaxies. The more clustered inactive galaxies are likely to be satellite galaxies in high-density environments (Fig.~\\ref{fig:mean-enviro} and Fig.~\\ref{fig:distrib-enviro}). \\end{itemize} The key to understanding this in more detail now lies in the dynamics of the central regions of galaxies, to understand what activates some galaxies, but not others." }, "0710/0710.4567_arXiv.txt": { "abstract": "We construct merger trees from the largest database of dark matter haloes to date provided by the Millennium simulation to quantify the merger rates of haloes over a broad range of descendant halo mass ($10^{12} \\la M_0 \\la 10^{15} M_\\odot$), progenitor mass ratio ($10^{-3} \\la \\xi \\le 1$), and redshift ($0 \\le z \\la 6$). We find the mean merger rate {\\it per halo}, $B/n$, to have very simple dependence on $M_0$, $\\xi$, and $z$, and propose a universal fitting form for $B/n$ that is accurate to 10-20\\%. Overall, $B/n$ depends very weakly on the halo mass ($\\propto M_0^{0.08}$) and scales as a power law in the progenitor mass ratio ($\\propto \\xi^{-2}$) for minor mergers ($\\xi \\la 0.1$) with a mild upturn for major mergers. As a function of time, we find the merger rate per Gyr to evolve roughly as $(1+z)^{n_m}$ with $n_m=2-2.3$, while the rate per unit redshift is nearly independent of $z$. Several tests are performed to assess how our merger rates are affected by, e.g. the time interval between Millennium outputs, binary vs multiple progenitor mergers, and mass conservation and diffuse accretion during mergers. In particular, we find halo fragmentations to be a general issue in merger tree construction from $N$-body simulations and compare two methods for handling these events. We compare our results with predictions of two analytical models for halo mergers based on the Extended Press-Schechter (EPS) model and the coagulation theory. We find that the EPS model overpredicts the major merger rates and underpredicts the minor merger rates by up to a factor of a few. ", "introduction": "In hierarchical cosmological models such as $\\Lambda$CDM, galaxies' host dark matter haloes grow in mass and size primarily through mergers with other haloes. As the haloes merge, their more centrally concentrated baryonic components sink through dynamical friction and merge subsequently. The growth of stellar masses depends on both the amount of mass brought in by mergers and the star formation rates. Having an accurate description of the mergers of dark matter haloes is therefore a key first step in quantifying the mergers of galaxies and in understanding galaxy formation and growth. Earlier theoretical studies of galaxy formation typically relied on merger trees generated from Monte Carlo realisations of the merger rates given by the analytical extended Press-Schechter (EPS; \\citealt{1993MNRAS.262..627L, 1991ApJ...379..440B}) model (e.g. \\citealt{1993MNRAS.264..201K, 1999MNRAS.310.1087S, 2000MNRAS.319..168C}). Some recent studies have chosen to bypass the uncertainties and inconsistencies in the EPS model by using halo merger trees from $N$-body simulations directly (\\citealt{1999MNRAS.303..188K, 2000MNRAS.311..793B, 2003MNRAS.338..903H, 2005ApJ...631...21K, 2005Natur.435..629S}). As we find in this paper, obtaining robust halo merger rates and merger trees requires rich halo statistics from very large cosmological simulations as well as careful treatments of systematic effects due to different algorithms used for, e.g., assigning halo masses, constructing merger trees, removing halo fragmentation events, and choosing time spacings between simulation outputs. The aim of this paper is to determine the merger rates of dark matter haloes as a function of halo mass, merger mass ratio (i.e. minor vs major), and redshift, using numerical simulations of the $\\Lambda$CDM cosmology. This basic quantity has not been thoroughly investigated until now mainly because large catalogues of haloes from finely spaced simulation outputs are required to provide sufficient merger event statistics for a reliable construction of merger trees over a wide dynamic range in time and mass. We achieve this goal by using the public database of the Millennium simulation \\citep{2005Natur.435..629S}, which follows the evolution of roughly $2\\times10^{7}$ dark matter haloes from redshift $z=127$ to $z=0$. This dataset allows us to determine the merger rates of dark matter haloes ranging from galaxy-mass scales of $\\sim 10^{12} M_\\odot$ over redshifts $z=0$ to $\\sim 6$, to cluster-mass scales up to $\\sim 10^{15} M_\\odot$ for $z=0$ to a few. We are also able to quantify the merger rates as a function of the progenitor mass ratio $\\xi$, from major mergers ($\\xi \\ga 0.1$) down to minor mergers of $\\xi \\sim 0.03$ for galaxy haloes and down to $\\xi \\sim 3\\times 10^{-4}$ for cluster haloes. The inputs needed for measuring merger rates in simulations include a catalogue of dark matter haloes and their masses at each redshift, and detailed information about their ancestry across redshifts, that is, the merger tree. Unfortunately there is not a unique way to identify haloes, assign halo masses, and construct merger trees. In this paper we primarily consider a halo mass definition based on the standard friends-of-friends (FOF) algorithm and briefly compare it with an alternative mass definition based on spherical overdensity. For the merger trees, we investigate two possible algorithms for treating events in which the particles in a given progenitor halo end up in more than one descendant halo ('fragmentations'). We find that these events are common enough that a careful treatment is needed. In the conventional algorithm used in the literature, the progenitor halo is linked one-to-one to the descendant halo that has inherited the largest number of the progenitor's particles. The ancestry links to the other descendant haloes are severed (for this reason we call this scheme 'snipping'). We consider an alternative algorithm ('stitching') in this paper, in which fragmentations are assumed to be artefacts of the FOF halo identification scheme. We therefore choose to recombine the halo fragments and stitch them back into the original FOF halo. Earlier theoretical papers on merger rates either relied on a small sample of main haloes to estimate the overall redshift evolution over a limited range of halo masses, or were primarily concerned with the mergers of {\\it galaxies} or {\\it subhaloes}. For halo mergers, for example, \\cite{1999AJ....117.1651G} studied $z < 1$ major mergers of galaxy-sized haloes in an open CDM and a tilted $\\Omega_m=1$ CDM model using $N$-body simulations in a 100 Mpc box and $144^3$ particles. \\cite{2001ApJ...546..223G} used a sample of $\\sim 4000$ haloes to study the environmental dependence of the redshift evolution of the major merger rate at $z < 2$ in $\\Lambda$CDM. \\cite{2006ApJ...652...56B} studied major mergers of subhaloes in $N$-body simulations in a 171 Mpc box with $512^3$ particles and the connection to the observed close pair counts of galaxies. For galaxy merger rates, \\cite{2002ApJ...571....1M} and \\cite{2006ApJ...647..763M} are based on up to $\\sim 500$ galaxies formed in SPH simulations in $\\sim 50$ Mpc boxes with up to $144^3$ gas particles, while \\cite{2007arXiv0708.1814G} used the semi-analytical galaxy catalogue of \\cite{2006MNRAS.366..499D} based on the Millennium simulation. This paper is organised as follows. Section~\\ref{MillenniumSection} describes the dark matter haloes in the Millennium simulation (\\S\\ref{DarkMatterHaloes}) and how we construct the merger trees (\\S\\ref{mergertrees}) . We then discuss the issue of halo fragmentation and the two methods ('snipping' and 'stitching') used to treat these events in \\S\\ref{FOFLims}. The notation used in this paper is summarised in \\S\\ref{Notation}. Section~\\ref{StatsSection} describes how mergers are counted (\\S\\ref{MultiCount}) and presents four (related) statistical measures of the merger rate (\\S\\ref{stats}). The relation between these merger rate statistics and the analytical merger rate based on the Extended Press-Schechter (EPS) model is derived in Section~\\ref{EPSConnection}. Our main results on the merger rates computed from the Millennium simulation are presented in Section~\\ref{Results}. We first discuss the $z\\approx 0$ results and quantify the merger rates as a function of the descendant halo mass and the progenitor mass ratios using merger trees constructed from the stitching method (\\S\\ref{MRz0}). The evolution of the merger rates with redshifts up to $z\\sim 6$ is discussed in Section~\\ref{redshiftdependence}. We find a simple universal form for the merger rates and present an analytic fitting form that provides a good approximation (at the 10-20\\% level) over a wide range of parameters (\\S\\ref{FitSection}). Section~\\ref{Tests} compares the stitching and snipping merger rates (\\S\\ref{SnipvsStitch}) and presents the key results from a number of tests that we have carried out to assess the robustness of our results. Among the tests are: time convergence and the dependence of the merger rates on the redshift spacing $\\dz$ between the Millennium outputs used to construct the merger tree (\\S\\ref{TimeResolutionSection}); how the counting of binary vs multiple progenitor mergers affects the merger rates (\\S\\ref{MultiCountValid}); mass non-conservation arising from 'diffuse' accretion in the form of unresolved haloes during mergers (\\S\\ref{MassCons}); and how the definition of halo masses and the treatment of fragmentation events affect the resulting halo mass function (\\S\\ref{MassFunction}). In Section~\\ref{DiscussionSection}, we discuss two theoretical frameworks that can be used to model halo mergers: EPS and coagulation. A direct comparison of our merger rates and the EPS predictions for the Millennium $\\Lambda$CDM model shows significant differences over a large range of parameter space (\\S\\ref{eps}). Section~\\ref{Coagulation} discusses Smoluchowski's coagulation equation and the connection between our merger rates and the coagulation merger kernel. The appendix compares a third merger tree (besides snipping and stitching) constructed from the Millennium catalogue by the Durham group \\citep{2006MNRAS.370..645B, 2006MNRAS.367.1039H,2003MNRAS.338..903H}. Two additional criteria are imposed on the subhaloes in this algorithm to reduce spurious linkings of FOF haloes. We find these criteria to result in reductions in both the major merger rates and the halo mass function. The cosmology used throughout this paper is identical to that used in the Millennium simulation: a $\\Lambda\\textrm{CDM}$ model with $\\Omega_m=0.25$, $\\Omega_b=0.045$, $\\Omega_\\Lambda=0.75$, $h=0.73$, an initial power-law index $n=1$, and $\\sigma_8=0.9$ \\citep{2005Natur.435..629S}. Masses and lengths are quoted in units of $M_\\odot$ and Mpc without the Hubble parameter $h$. ", "conclusions": "In this paper we have computed the merger rates of dark matter FOF haloes as a function of descendant halo mass $M_0$, progenitor mass ratio $\\xi$, and redshift $z$ using the merger trees that we constructed from the halo catalogue of the Millennium simulation. Our main results are presented in Figs.~\\ref{fig:B} to \\ref{fig:Rz}, which show very simple and nearly separable dependence on $M_0$, $\\xi$, and $z$. The mean merger rate per descendant FOF halo, $B/n$, is seen to depend very weakly on the halo mass $M_0$ (Fig.~\\ref{fig:B} right panel and Fig.~\\ref{fig:Rmass}). As a function of redshift $z$, the per halo merger rate in units of per Gyr increases as $(1+z)^\\alpha$, where $\\alpha\\sim 2$ to 2.3 (top panel of Fig.~\\ref{fig:Rz}), but when expressed in units of per redshift, the merger rate depends very weakly on $z$ (bottom panel of Fig.~\\ref{fig:Rz}). Regardless of $M_0$ and $z$, the dependence of $B/n$ on the progenitor mass ratio, $\\xi = M_i/M_1$, is a power law to a good approximation in the minor merger regime ($\\xi \\la 0.1$) and shows an upturn in the major merger regime (Fig.~\\ref{fig:B}). These simple behaviours have allowed us to propose a universal fitting formula in equation~(\\ref{fiteqn}) that is valid for $10^{12}\\leq M_0\\la 10^{15} M_\\odot$, $\\xi \\ga 10^{-3}$, and up to $z\\sim 6$. Throughout the paper we have emphasised and quantified the effects on the merger rates due to events in which a progenitor halo fragments into multiple descendant haloes. We have shown that the method commonly used to remove these fragmented haloes in merger trees -- the snipping method -- has relatively poor $\\dz$-convergence (Figs.~\\ref{fig:TR} and \\ref{fig:RzTR}). Our alternative approach -- the stitching method -- performs well with regards to this issue without drastically modifying the mass conservation properties or the mass function of the Millennium FOF catalogue (Figs.~ \\ref{fig:DM} and \\ref{fig:MassFunction}). We have computed the two predictions for merger rates from the analytical EPS model for the same $\\Lambda$CDM model used in the Millennium simulation. At $z=0$, we find the EPS major merger rates to be too high by 50-100\\% (depending on halo mass) and the minor merger rates to be too low by up to a factor of 2-5 (Fig.~\\ref{fig:EPS}). The discrepancy increases at higher $z$. The coagulation equation offers an alternative theoretical framework for modelling the mergers of dark matter haloes. We have discussed how our merger rate is related to the coagulation merger kernel in theory. In practice, however, we find that mergers in simulations are not always mass-conserving binary events, as assumed in the standard coagulation form given by equation~(\\ref{eqn:coag}). Equation~(\\ref{eqn:coag}) will therefore have to be modified before it can be used to model mergers in simulations. \\cite{2001ApJ...546..223G} studied the rate of major mergers (defined to be $\\xi \\ge 1/3$ in our notation) in $N$-body simulations and found a steeper power law dependence of $\\propto (1+z)^3$ (at $z\\la 2$) for the merger rate per Gyr than ours. Their simulations did not have sufficient mass resolution to determine the rate at $z \\ga 2$. It is important to note, however, that our $B/n$ at redshift $z$ measures the instantaneous rate of mergers during a small $\\Delta z$ interval at that redshift. By contrast, they studied the merging history of {\\it present-day} haloes and measured only the rate of major mergers for the most massive progenitor at redshift $z$ of a $z=0$ halo (see their paragraph 4, section 2). A detailed comparison is outside the interest of this paper. Mergers of dark matter haloes are related to but not identical to mergers of galaxies. It typically takes the stellar component of an infalling galaxy extra time to merge with a central galaxy in a group or cluster after their respective dark matter haloes have been tagged as merged by the FOF algorithm. This time delay is governed by the dynamical friction timescale for the galaxies to lose orbital energy and momentum, and it depends on the mass ratios of the galaxies and the orbital parameters (\\citealt{2008MNRAS.383...93B} and references therein). In addition to this difference in merger timescale, the growth in the stellar mass of a galaxy is not always proportional to the growth in its dark matter halo mass. A recent analysis of the galaxy catalogue in the Millennium simulation \\citep{2007arXiv0708.1814G} finds galaxy growth via major mergers to depend strongly on stellar mass, where mergers are more important in the buildup of stellar masses in massive galaxies while star formation is more important in galaxies smaller than the Milky Way. Extending the analysis of this paper to the mergers of {\\it subhaloes} in the Millennium simulation will provide the essential link between their and our results. For similar reasons, our results for the evolution of the dark matter halo merger rate per Gyr ($(1+z)^{n_m}$ with $n_m\\sim2-2.3$) cannot be trivially connected to the observed merger rate of \\emph{galaxies}. It is nonetheless interesting to note that a broad disagreement persists in the observational literature of galaxy merger rates. The reported power law indices $n_m$ have ranged from 0 to 5 (see, for example, \\citealt{1994ApJ...429L..13B, 1994ApJ...435..540C, 1995ApJ...445...37Y, 1995ApJ...454...32W, 1997ApJ...475...29P, 2000MNRAS.311..565L, 2002ApJ...565..208P,2003AJ....126.1183C, 2004ApJ...601L.123B,2004ApJ...612..679L, 2004ApJ...617L...9L}). \\cite{2006ApJ...652...56B} followed the redshift evolution of subhalo mergers in N-body simulations and provided a more detailed comparison with recent observations by, e.g., \\cite{2004ApJ...617L...9L} that find $n_m<1$. They attributed such a weak redshift evolution in the number of close companions per galaxy to the fact that the high merger rate per halo at early times is counteracted by a decrease in the number of haloes massive enough to host a galaxy pair. The merger rates in this paper are global averages over all halo environments. The rich statistics in the Millennium simulation allow for an in-depth analysis of the environmental dependence of dark matter halo merger rates, which we will report in a subsequent paper (Fakhouri \\& Ma, in preparation)." }, "0710/0710.1945_arXiv.txt": { "abstract": "{The ARGO-YBJ experiment is a full coverage EAS-array installed at the YangBaJing Cosmic Ray Laboratory (4300 m a.s.l., Tibet, P.R. China). We present the results on the angular resolution measured with different methods with the full central carpet. The comparison of experimental results with MC simulations is discussed.} \\begin{document} ", "introduction": "The ARGO-YBJ detector is constituted by a single layer of Resistive Plate Chambers (RPCs). This carpet has a modular structure, the basic unit is a cluster, composed by 12 RPCs (2.8$\\times$1.25 m$^2$ each). Each chamber is read by 80 strips, logically organized in 10 independent pads\\cite{nim_argo}. The central carpet, constituted by 10$\\times$13 clusters with $\\sim$93$\\%$ of active area, is enclosed by a guard-ring partially instrumented ($\\sim$40$\\%$) in order to improve rejection capability for external events. A lead converter 0.5 cm thick will uniformly cover the apparatus in order to improve the angular resolution. Since December 2004 the pointing accuracy of the detector has been studied, during the detector setting-up, with 3 different carpet areas: 42 clusters (ARGO-42, $\\sim$1900 m$^2$), 104 clusters (ARGO-104, $\\sim$4600 m$^2$) and the full central carpet, 130 clusters (ARGO-130, $\\sim$5800 m$^2$), yet without any converter sheet. The data have been collected with a so-called {\\it \"Low Multiplicity Trigger\"}, requiring at least $20$ fired pads on the whole detector. ", "conclusions": "Since December 2004 increasing fractions of ARGO-YBJ detector have been put in data taking even with a reduced duty-cycle due to installation and debugging operations. In this paper we presented a measurement of the pointing accuracy of the ARGO-130 detector. The capability of reconstructing the primary shower direction has been investigated with the chessboard method and with a preliminary Moon shadow analysis. Studies are in progress in order to determine the final angular resolution." }, "0710/0710.5427_arXiv.txt": { "abstract": "\\footnotesize\\ The efficient numerical solution of Non-LTE multilevel transfer problems requires the combination of highly convergent iterative schemes with fast and accurate formal solution methods of the radiative transfer (RT) equation. This contribution\\footnote{Published in 1999 in the book {\\it Solar Polarization}, edited by K.N. Nagendra \\& J.O. Stenflo. Kluwer Academic Publishers, 1999. (Astrophysics and Space Science Library ; Vol. 243), p. 219-230} begins presenting a method for the formal solution of the RT equation in three-dimensional (3D) media with horizontal periodic boundary conditions. This formal solver is suitable for both, unpolarized and polarized 3D radiative transfer and it can be easily combined with the iterative schemes for solving non-LTE multilevel transfer problems that we have developed over the last few years. We demonstrate this by showing some schematic 3D multilevel calculations that illustrate the physical effects of horizontal radiative transfer. These Non-LTE calculations have been carried out with our code MUGA 3D, a 3D multilevel Non-LTE code based on the Gauss-Seidel iterative scheme that Trujillo Bueno and Fabiani Bendicho (1995) developed for RT applications. ", "introduction": "To what extent can we trust diagnostic results obtained with the assumption that the solar atmospheric plasma is composed of {\\it homogeneous} plane-parallel layers or via approximations that neglect {\\it horizontal} radiative transfer (RT) effects? How important are the errors in the magnetic fields, temperatures and velocities inferred by confronting spectro-polarimetric observations with Non-LTE 1D RT model calculations? Clearly, to provide proper answers to questions like these requires to develop first efficient 3D RT methods that allow Non-LTE effects in complex atomic models with many levels and transitions to be rigorously investigated. There is a second reason which makes the development of fast iterative methods for 3D Non-LTE RT so relevant. This is because processes of energy exchange by radiation play an important role in the structure and dynamical behaviour of the stellar magnetized plasma. Thus, for instance, if one wishes to perform time-dependent radiation hydrodynamics simulations similar to those carried out by Carlsson and Stein (1997), but in 3D instead of 1D, it turns out to be imperative to have first access to numerical methods capable of accurately yielding the self-consistent atomic level populations at the cost of only {\\it very} few formal solution times. The efficient solution of multilevel transfer problems requires the combination of a highly convergent iterative scheme with a fast formal solver of the RT equation. In Section 2 we briefly comment on a hierarchy of iterative schemes that can be applied for solving multilevel Non-LTE problems with increasing improvements in the convergence rate and total computational work. The 3D multilevel transfer calculations that we present in this contribution have been obtained by combining a highly convergent iterative scheme based on Gauss-Seidel iteration (Trujillo Bueno and Fabiani Bendicho, 1995) with a fast 3D formal solver that has parabolic accuracy (see Section 3). As was the case with our 2D formal solver, our generalization to 3D is based on the ``short-characteristics'' method of Kunasz and Auer (1988). Our 3D multilevel code is called MUGA 3D (``Multi-level Gauss-Seidel Method'') and it is substantially {\\it faster} than our code MALI 3D, which is based on Jacobi iteration (see Rybicki and Hummer, 1991; Auer, Fabiani Bendicho and Trujillo Bueno, 1994). Section 3 briefly describes our 3D formal solver as applied to the scalar transfer equation for the specific intensity (I). In order to be able to consider 3D atmospheric models where solar plasma structures repeat themselves along the horizontal directions we choose horizontal periodic boundary conditions along the Cartesian coordinates X and Y. Although we do not give any details here, we have also generalized to 3D the Stokes-vector 1D formal solver method developed by Trujillo Bueno (1998), which is based on the matrix exponential approximation to the evolution operator. In Section 4 we show some illustrative 3D multilevel transfer calculations for a 5-level Ca II model atom where the H, K and infrared triplet lines are treated simultaneously, taking fully into account the {\\it interlockings} by which photons are converted back and forth between the different line transitions in the assumed 3D medium. Here we consider schematic 3D solar models characterized by horizontal sinusoidal temperature inhomogeneities. With the help of these 3D multilevel calculations we are able to illustrate some subtle effects of horizontal radiative transfer that are important for the correct interpretation of high spatial resolution observations. Finally, Section 5 gives our conclusions. ", "conclusions": "We have developed a 3D multilevel code (MUGA-3D) that combines the Gauss-Seidel iterative scheme of Trujillo Bueno and Fabiani Bendicho (1995) with a 3D formal solver that uses horizontal periodic boundary conditions. With this new code we have performed some 3D multilevel simulations that highlight the importance of carefully investigating the effects of horizontal radiative transfer using realistic atmospheric and atomic models. We point out that our 3D formal solver can be used not only for solving ``unpolarized'' multilevel transfer problems, but also resonance line polarization and Hanle effect problems, like those considered in these Proceedings by Manso Sainz and Trujillo Bueno (1999), Paletou {\\it et. al.} (1999) or Dittmann (1999). This is because, for these polarization transfer cases, the absorption matrix is diagonal. As a result, we have similar equations for the Stokes I, Q and U parameters. However, for the solution of more general polarization transfer problems, like the Non-LTE Zeeman line transfer case considered by Trujillo Bueno and Landi Degl'Innocenti (1996), but in 3D instead of 1D, one needs a 3D formal solution method of the Stokes-vector transfer equation. This is because here the absorption matrix turns out to be a full $4\\times4$ matrix and all the Stokes parameters are coupled together. To this end we have generalized to 3D the Stokes-vector formal solver developed by Trujillo Bueno (1998), which can be considered as a generalization to polarization transfer of the short-characteristics method. We would like to end this paper by saying that over the last 10 years we have witnessed impressive progress concerning the development of highly convergent iterative schemes and accurate formal solvers for RT applications. Now it is time to apply them with physical intuition in order to improve our knowledge of the Sun, its magnetic field and its polarized spectrum." }, "0710/0710.1338_arXiv.txt": { "abstract": "The two-body problem in general relativity is reviewed, focusing on the final stages of the coalescence of the black holes as uncovered by recent successes in numerical solution of the field equations. ", "introduction": "A black hole is one of the most fascinating and enigmatic predictions of Einstein's theory of general relativity. Its interior can have rich structure and is intrinsically dynamical, where space and time itself are inexorably led to a singular state. The exterior of an isolated black hole is, on the other hand, remarkably simple, described uniquely by the stationary Kerr solution. The dynamics of black holes are governed by laws analogous to the laws of thermodynamics, and indeed when quantum processes are included, emit Hawking radiation with a characteristic thermal spectrum. Most remarkable however, is that black holes, ``discovered'' purely through thought and the mathematical exploration of a theory far removed from every day experience, appear to be ubiquitous objects in our universe. The evidence that black holes exist, though circumstantial, is quite strong~\\cite{Narayan:2005ie}. The high luminosity of quasars and other active galactic nuclei (AGN) can be explain by gravitational binding energy released through gas accretion onto supermassive ($10^6-10^9 \\msun$) black holes at the centers of the galaxies~\\cite{Rees:1984si,Ferrarese:2004qr}, several dozen X-ray binary systems discovered to date have compact members too massive to be neutron stars and exhibit phenomena consistent with matter interactions originating in the strong gravity regime of an inner accretion disk~\\cite{McClintock:2003gx}, and the dynamical motion of stars and gas about the centers of nearby galaxies and our Milky Way Galaxy infer the presence of very massive, compact objects there, the most plausible explanation being supermassive black holes~\\cite{Gebhardt:2000fk,Schodel:2002py,Ghez:2003qj}. To conclusively prove that black holes exist one needs to ``see'' them, or conversely see the compact objects masquerading as black holes. The only direct way of observing black holes is via the gravitational waves they emit when interacting with other matter/energy (an isolated black hole does not radiate). The quadrupole formula says that the typical magnitude $h$ of the gravitational waves emitted by a binary with reduced mass $\\mu$ on a circular orbit measured a distance $r$ from the source is (for a review of gravitational wave theory see~\\cite{Flanagan:2005yc}) \\be h=\\frac{16 \\mu v^2}{r}, \\ee where $v$ is the average tangential speed of the two members in the binary (and geometric units are used---Newton's constant $G=1$ and the speed of light $c=1$). This formula suggests that the strongest sources of gravitational waves are simply the most massive objects that move the fastest. To reach large velocities in orbit, the binary separation has to be small; black holes, being the most compact objects allowed in the theory, can reach the closest possible separations and hence largest orbital velocities. Therefore, modulo questions about source populations in the universe, a binary black hole interaction offers one of most promising venues of observing black holes through gravitational wave emission. Joseph Weber pioneered the science of gravitational wave detection with the construction of resonant bar detectors. Weber claimed to have detected gravitational waves~\\cite{Weber:1969bz}, though no similar detectors constructed following his claims were able to observe the putative (or any other) source, and the general consensus is that given the sensitivity of Weber's detector and expected strengths of sources it is very unlikely that it was a true detection~\\cite{Thorne:1980rt}. Note that the {\\em existence} of gravitational waves is not in doubt---the observed spin down rate of the Hulse-Taylor binary pulsar~\\cite{Hulse:1974eb} and several others discovered since, is in complete accord with the general relativistic prediction of spin down via gravitational wave emission. Today a new generation of gravitational wave detectors are operational, including laser interferometers (LIGO~\\cite{LIGO}, VIRGO~\\cite{VIRGO}, GEO600~\\cite{GEO}, TAMA~\\cite{TAMA}) and resonant bar detectors (NAUTILUS~\\cite{NAUTILUS}, EXPLORER~\\cite{EXPLORER}, AURIGA~\\cite{AURIGA}, ALLEGRO~\\cite{ALLEGRO}, NIOBE~\\cite{NIOBE}). A future space-based observatory is planned (LISA~\\cite{LISA}), and pulsar timing and cosmic microwave background polarization measurements also offer the promise of acting as gravitational wave ``detectors'' (for reviews see~\\cite{Maggiore:1999vm,Cutler:2002me}). The ultimate success of gravitational wave detectors, in particular with regards to using them as more that simply detectors, but tools to observe and understand the universe, relies on source modeling to predict the structure of the waves emitted during some event. Even if an event is detected with a high signal-to-noise ratio (SNR), there simply is not enough information contained in such a one dimensional time series to ``invert'' it to reconstruct the event; rather template banks of theoretical waveforms from plausible sources need to be built and used to decode the signal. In rare cases an electromagnetic counterpart may be detected, for example during a binary neutron star merger if this is a source of short gamma ray bursts, which could identify the event without the need for a template. Though even in such a case, to extract information about the event, its environment, etc. requires source modeling. Gravitational wave detectors have therefore provided much of the impetus for trying to understand the nature of binary black hole collisions, and the gravitational waves emitted during the process. However, from a theoretical perspective black hole collisions are fantastic probes of the dynamical, strong-field regime of general relativity. What is already know about this regime---the inevitability of spacetime singularities in gravitational collapse via the singularity theorems of Penrose and Hawking~\\cite{Hawking:1969sw,hawking_ellis}; the spacelike, chaotic ``mixmaster'' nature of these singularities conjectured by Belinsky, Khalatnikov and Lifshitz (BKL)~\\cite{Berger:1998us}; the null, mass-inflation singularity discovered by Poisson and Israel~\\cite{Poisson:1990eh} that, together with regions of BKL singularities could generically describe the interiors of black hole; the rather surprising discovery of critical phenomena in gravitational collapse by Choptuik~\\cite{Choptuik:1992jv,Gundlach:1999cu}; etc---together with the sparsity of solutions (exact, numerical or perturbative), suggests there is potentially a vast landscape of undiscovered phenomena. Of particular interest, and potential application to high energy particle collision experiments, are ultra-relativistic black hole collisions. It is beyond the scope of this article to delve much into these aspects of black hole coalescence, though a brief overview of this will be given in Sec.~\\ref{sec_he}. The two body problem in general relativity, introduced in more detail in Sec.~\\ref{sec_2bdy}, is a very rich and complicated problem, with no known closed-form solution. Perturbative analytic techniques have been developed to deal with certain stages of the problem, in particular the inspiral prior to merger and ringdown after merger. Numerical solution of the full field equations are required during the merger, and this aspect of the problem is the main focus of this article. Much effort has been expended by the community over the past 15-20 years to numerically solve for merger spacetimes, and within the last two years an understanding of this phase of the two body problem is finally being attained. Sec.~\\ref{sec_num} summaries the difficulties in discretizing the field equations, and describes the methods known at present that work for black hole collisions, namely {\\em generalized harmonic coordinates} and {\\em BSSN} (Baumgarte-Shapiro-Shibata-Nakamura) with moving punctures. Preliminary results are discussed in Sec.~\\ref{sec_res}, though given the rapid pace at which the field is developing much of this will probably be dated in short order. Sec.~\\ref{sec_imp} concludes with a discussion of some astrophysical and other implications of the results. ", "conclusions": "The two body problem in general relativity is a fascinating, rich problem that is just beginning to be fully revealed by recent breakthroughs in numerical relativity. At the same time, a new generation of gravitational wave detectors promise to offer us a view of the universe via the gravitational wave spectrum for the first time. Black hole mergers are a promising source for gravitational waves, and detecting them would provide direct evidence for these remarkable objects, while providing much information about their environments. Suggestions that there might be more than four spacetime dimensions offers the astonishing possibility that black holes could be produced by proton collisions at $TeV$ energies, which will be reached by the Large Hadron Collider, planned to begin operation within a year. Given all this, it is difficult not to be excited about what might be learnt about the universe from the smallest to largest scales during the next decade, and that black hole collisions could have something important to say at both extremes. \\bigskip \\noindent{\\bf{\\em Acknowledgments:}} I would like to thank Alessandra Buonanno, Matthew Choptuik, Gregory Cook, Charles Gammie, David Garfinkle, Steven Gubser, Carsten Gundlach, Luis Lehner, Jeremiah Ostriker, Don Page, David Spergel and Ulrich Sperhake for many stimulating conversations related to some of the discussion presented here." }, "0710/0710.5388_arXiv.txt": { "abstract": "Recently, it has been shown that the standard Nambu-Jona-Lasinio (NJL) model is not able to reproduce the correct QCD behavior of the gap equation at large density, and therefore a different cutoff procedure at large momenta has ben proposed. We found that, even with this density dependent cutoff procedure, the pure quark phase in neutron stars (NS) interiors is unstable, and we argue that this could be related to the lack of confinement in the original NJL model. ", "introduction": " ", "conclusions": "" }, "0710/0710.5207_arXiv.txt": { "abstract": "If massive black holes (BHs) are ubiquitous in galaxies and galaxies experience multiple mergers during their cosmic assembly, then BH binaries should be common albeit temporary features of most galactic bulges. Observationally, the paucity of active BH pairs points toward binary lifetimes far shorter than the Hubble time, indicating rapid inspiral of the BHs down to the domain where gravitational waves lead to their coalescence. Here, we review a series of studies on the dynamics of massive BHs in gas-rich galaxy mergers that underscore the vital role played by a cool, gaseous component in promoting the {\\it rapid formation of the BH binary}. The BH binary is found to reside at the center of a massive self-gravitating nuclear disc resulting from the collision of the two gaseous discs present in the mother galaxies. Hardening by gravitational torques against gas in this grand disc is found to continue down to sub-parsec scales. The eccentricity decreases with time to zero and when the binary is circular, accretion sets in around the two BHs. When this occurs, each BH is endowed with it own small-size ($\\simlt 0.01$ pc) accretion disc comprising a few percent of the BH mass. Double AGN activity is expected to occur on an estimated timescale of $\\simlt 1$ Myr. The double nuclear point--like sources that may appear have typical separation of $\\simlt 10$ pc, and are likely to be embedded in the still ongoing starburst. We note that a potential threat of binary stalling, in a gaseous environment, may come from radiation and/or mechanical energy injections by the BHs. Only short--lived or sub--Eddington accretion episodes can guarantee the persistence of a dense cool gas structure around the binary necessary for continuing BH inspiral. ", "introduction": "Dormant black holes (BHs) with masses in excess of $\\simgt 10^6\\msun$ are ubiquitous in bright galaxies today (Kormendy \\& Richstone 1995; Richstone 1998). Relic of an earlier active phase as quasars, these massive BHs appear a clear manifestation of the cosmic assembly of galaxies. The striking correlations observed between the BH masses and properties of the underlying hosts (Magorrian \\etal 1998; Ferrarese \\& Merritt 2000; Gebhardt et al. 2000; Graham \\& Driver 2007) indicates unambiguously that BHs evolve in symbiosis with galaxies, affecting the environment on large--scales and self-regulating their growth (Silk \\& Rees 1998; King 2000; Granato et al. 2004; Di Matteo, Springel \\& Hernquist, 2005). According to the current paradigm of structure formation, galaxies often interact and collide as their dark matter halos assemble in a hierarchical fashion (Springel, Frenk \\& White 2006), and BHs incorporated through mergers into larger and larger systems are expected to evolve concordantly (Volonteri, Haardt \\& Madau 2003). In this astrophysical context, close BH {\\it pairs} form as natural outcome of binary galaxy mergers (Kazantzidis et al. 2005). In our local universe, one outstanding example is the case of the ultra--luminous infrared galaxy NGC 6240, an ongoing merger between two gas--rich galaxies (Komossa \\etal 2003; for a review on binary black holes see also Komossa 2006). {\\it Chandra} images have revealed the occurrence of two nuclear X-ray sources, 1.4 kpc apart, whose spectral distribution is consistent with being two accreting massive BHs embedded in the diluted emission of a starburst. Similarly, Arp 299 (Della Ceca et al. 2002; Ballo et al. 2004) is an interacting system hosting an obscured active nucleus, and possibly a second less luminous one, distant several kpc away. A third example is the elliptical galaxy 0402+369 where the cleanest case of a massive BH {\\it binary} has been recently discovered. Two compact variable, flat--spectrum active nuclei are seen at a projected separation of only 7.3 pc (Rodriguez et al. 2006). Arp 299, NGC 6240, and 0402+369 may just highlight different stages of the BH dynamical evolution along the course of a merger, with 0402+369 being the latest, most evolved phase (possibly related to a dry merger). Energy and angular momentum losses due to gravitational waves are not yet significant in 0402+392, so that stellar interactions and/or material and gravitational torques are still necessary to bring the BHs down to the domain controlled by General Relativity. From the above considerations and observational findings, it is clear that binary BH inspiral down to coalescence is a major astrophysical process that can occur in galaxies. It is accompanied by a gravitational wave burst so powerful to be detectable out to very large redshifts with current planned experiments like the Laser Interferometer Space Antenna ({\\it LISA}; Bender et al. 1994; Vitale et al. 2002). These extraordinary events will provide not only a firm test of General Relativity but also a view, albeit indirect, of galaxy clustering (Haehnelt 1994; Jaffe \\& Backer 2003; Sesana et al. 2005). With {\\it LISA}, BH masses and spins will be measured with such an accuracy (Vecchio 2004) that it will be possible to trace the BH mass growth across all epochs. Interestingly, {\\it LISA} will explore a mass range between $10^3\\msun$ and $10^7\\msun$ that is complementary to that probed by the distant massive quasars ($>10^7\\msun$), providing a complete census of the BHs in the universe. Both minor as well as major mergers with BHs accompany galaxy evolution in environments that involve either gas-rich (wet) as well as gas-poor (dry) galaxies. Thus, the dynamical response of galaxies to BH pairing should differ in many ways according to their properties. Exploring the expected diversities in a self-consistent cosmological scenario is a major challenge and only recently, with the help of high-resolution N-body/SPH simulations, it has become possible to ``start'' addressing a number of compelling issues. Galaxy mergers cover cosmological volumes (a few to hundred kpc aside), whereas BH mergers probe volumes of only few astronomical units or less. Thus, tracing the BH dynamics with scrutiny requires N-Body/SPH force resolution simulations spanning more than nine orders of magnitude in length. For this reason, two complementary approaches have been followed in the literature. A statistical approach (based either on Monte Carlo realizations of merger trees or on N-Body/SPH large scale simulations) follows the collective growth of BHs inside dark matter halos. Supplemented by semi-analytical modeling of BH dynamics (Volonteri et al. 2003) or/and by sub-grid resolution criteria for accretion and feedback (Springel \\& Hernquist 2003; Springel, Di Matteo \\& Hernquist 2005), these studies have proved to be powerful in providing estimates of the expected coalescence rates, and in tracing the overall cosmic evolution of BHs including their feedback on the galactic environment (Di Matteo, Springel \\& Hernquist 2005; Di Matteo \\etal 2007). The second approach, that we have been following, looks at individual binary collisions, as it aims at exploiting in detail the BH dynamics and some bulk physics from the galactic scale down to and within the BH sphere of influence. Both approaches, the collective and the individual, are necessary and complementary, the main challenge being the implementation of realistic input physics in the dynamically active environment of a merger. Following a merger, how can BHs reach the gravitational wave inspiral regime? The overall scenario was first outlined by Begelman, Blandford \\& Rees (1980) in their seminal study on the dynamical evolution of BH pairs in pure stellar systems. They indicated three main roots for the loss of orbital energy and angular momentum: (I) dynamical friction against the stellar background acting on each individual BH; (II) hardening via 3--body scatterings off single stars when the BH binary forms; (III) gravitational wave back--reaction. Early studies explored phase (I) simulating the collisionless merger of spherical halos (Makino \\& Ebisuzaki 1996; Milosavljevi\\'c \\& Merritt 2001; Makino \\& Funato 2004). Governato, Colpi \\& Maraschi (1994) in particular first noticed that when two equal mass halos merge, the twin BHs nested inside the nuclei are dragged effectively toward the center of the remnant galaxy by dynamical friction and form a close pair, but that the situation reverses in unequal mass mergers, where the less massive halo tidally disrupted leaves its ``naked'' BH wandering in the outskirts of the main halo. Thus, depending on the halo mass ratio and internal structure, the transition from phase (I) to phase (II) can be prematurely aborted or drastically relented. Similarly, the transit from phase (II) to phase (III) is not always secured, as the stellar content inside the ``loss cone\" may not be rapidly refilled with fresh low--angular momentum stars to harden the binary down to separations where gravitational wave driven inspiral sets in (see, e.g., Milosavljevic \\& Merritt 2001; Yu 2002; Berczik, Merritt \\& Spurzem 2005; Sesana, Haardt \\& Madau 2007). For an updated review on the last parsec problem and its possible solution (see Merritt 2006a; Gualandris \\& Merritt 2007). Since BH coalescences are likely to be events associated with mergers of (pre--)galactic structures at high redshifts, it is likely that their dynamics occurred in gas dominated backgrounds, NGC 6240 being just the most outstanding case visible in our local universe. Other processes of BH binary hardening are expected to operate in presence of a dissipative gaseous component that we will highlight and study here. Kazantzidis \\etal (2005) first explored the effect of gaseous dissipation in mergers between gas--rich disc galaxies with central BHs, using high resolution N--Body/SPH simulations. They found that the merger triggers large--scale gas dynamical instabilities that lead to the gathering of cool gas deep in the potential well of the interacting galaxies. In minor mergers, this fact is essential in order to bring the BHs to closer and closer distances before the less massive galaxy, tidally disrupted, is incorporated in the main galaxy. Moreover, the interplay between strong gas inflows and star formation leads naturally to the formation, around the two BHs, of a grand, massive ($\\sim 10^9\\msun$) gaseous disc on a scale smaller than $\\sim 100$ pc. It is in this equilibrium circum--nuclear disc that the dynamical evolution of the BHs continues, after the merger has been subsided. Escala \\etal (2005, hereinafter ELCM05; see also Escala \\etal 2004) have been the first to study the role played by gas in affecting the dynamics of massive ($\\sim 10^8\\msun$) twin BHs in equilibrium Mestel discs of varying clumpiness. In both these approaches (i.e., in the large scale simulations of Kazantzidis et al., and in the equilibrium disc models of ELCM05) it was clear that the gas temperature is a key physical parameter and that a hot gas brakes the BHs inefficiently. Instead, when the gas is allowed to cool, the drag becomes efficient: the large enhancement of the local gas density relative to the stellar one leads to the formation of prominent density wakes that are decelerating the BHs down to the scale where they form a ``close'' binary. Later, binary hardening occurs under mechanisms that are only partially explored, and that are now subject of intense investigation. The presence of a cool circum--binary disc and of small--scale discs around each individual BH appear to be critical for their evolution down to the domain of gravitational waves driven inspiral. In this context there is no clear ``stalling problem'' that emerges from current hydrodynamical simulations but this critical phase need a more through, coherent analysis. The works by Kazantzidis \\etal (2005) and ELCM05 have provided our main motivation to study the process of BH pairing along two lines: In gas-rich binary mergers, line (1) aims at studying the transit from state (A) of pairing when each BH moves individually inside the time-varying potential of the colliding galaxies, to state (B) when the two BHs dynamically couple their motion to form a binary. The transit from (A) $\\to$ (B) requires exploring a dynamic range of five orders of magnitude in length from the cosmic scale of a galaxy merger of 100 kpc down to the parsec scale for BHs of million solar masses (i.e., BHs in the LISA sensitivity domain). After all transient inflows have subsided and a new galaxy has formed, the BH binary is expected to enter phase (C) where it hardens under the action of gas-dynamical and gravitational torques. Research line (2) aims at studying the braking of the BH binary from (B) $\\to$ (C) and further in, exploring the possibility that during phase (C) two discs form and grow around each individual BH. As first discussed by Gould \\& Rix (2000) the binary may later enter a new phase (D) controlled by the balance of viscous and gravitational torques in a circum--binary disc surrounding the BHs, in a manner analogous to the migration of planets in circum-stellar discs (a scenario particularly appealing when the BH mass ratio is less than unity). Phase (D) likely evolves into (E) when gravitational wave inspiral terminates the BH binary evolution. There is a number of key questions to address: \\noindent (i) How does transition from state (A) $\\to$ (B) depend on the gas thermodynamics? How do BHs bind? \\noindent (ii) In the grand nuclear disc inside the remnant galaxy, how do eccentric orbits evolve? Do they become circular or highly eccentric? \\noindent (iii) During the hardening through phase (B) and (C), do the BHs collect substantial amounts of gas to form cool individual discs? \\noindent (iv) Can viscous torques drive the binary into the gravitational wave decaying phase? \\noindent (v) Is there a threat of a {\\it stalling} problem when transiting from (C) $\\to$ (D) or from (D) $\\to$ (E)? And, for which mass ratios and ambient conditions? ", "conclusions": "" }, "0710/0710.0368_arXiv.txt": { "abstract": "We compute opacities for the electronic molecular band systems \\ACrH -- \\XCrH of CrH and CrD, and \\AMgH -- \\XMgH of MgH and MgD. The opacities are computed by making use of existing spectroscopic constants for MgH and CrH. These constants are adjusted for the different reduced masses of MgD and CrD. Frank-Condon factors are used to provide intensities for the individual vibronic bands. These results are used in the computation of synthetic spectra between \\Tef = 1800 and 1200 K with an emphasis on the realisation of ``deuterium test'', first proposed by Bejar et al. (1999) to distinguish brown dwarfs from planetary mass objects. We discuss the possible use of CrD and MgD electronic bands for the ``deuterium test\". We find CrD to be the more promising of the two deuterides, potentially, the most useful bands of CrH/CrD are the $\\Delta v = +1$ and $\\Delta v = -1$ at 0.795 and 0.968 \\mum . ", "introduction": "The ``deuterium test'' was suggested as a method of identifying planetary mass objects among cool objects (\\Bejar et al. 1999, Chabrier et al. 2000). In practise, it was proposed to search for absorption lines of molecules containing deuterium (HDO, CrD, FeD, etc.). Deuterium is burnt in stellar interiors via the fusion reaction $^2$D(p,$\\gamma$)$^3$He at temperatures ($T > 8 \\times 10^5$~ K). The interiors of substellar objects with M $<$ 13~ \\MJ, where \\MJ is a Jovian mass (0.001 \\Msun), do not reach temperatures high enough for deuterium to ignite (Saumon et al. 1996). As a result the deuterium abundance in the atmospheres of these objects is unchanged from the formation of these objects. This gives rise to the definition of a brown dwarf as an object which has insufficient mass to fuse $^1$H to $^4$He, but has sufficient mass to fuse $^2$H to $^3$He. By comparison, a planetary mass object has insufficient mass to ignite fusion of any sort (Saumon et al. 1996). In higher-mass objects such as stars, deuterium burning is completed comparatively quickly (t = 1 -- 3 million years) during the evolution prior to the star's period on the main sequence (D'Antona \\& Mazitelli 1998). The deuterium depletion rate depends on the mass of the star or brown dwarf and thus the ``deuterium test'' can be used in a number of ways. \\begin{itemize} \\item Discern planets from brown dwarfs in a population of low-mass objects. \\item Determine the evolutionary status of young objects in open clusters with ages of several million years. \\item Study the evolution of the abundance of deuterium in the atmospheres of low-mass substellar objects, a phenomenon which is poorly understood. The rate of depletion of deuterium depends upon rotation, magnetic field strength, and other parameters which affect the efficiency of convection in low mass objects. \\end{itemize} Such investigations can usefully be combined with the ``lithium test'' proposed by Rebolo et al. (1992), \\Magazzu et al. (1993). For stars (M $>$ 75 \\MJ), the burning of lithium, Li (p,$\\alpha) ^4$He, becomes efficient at evolutionary stages preceding the main sequence at interior temperatures of T $>$ $2.5\\times 10^6$ K (D'Antona \\& Mazitelli 1998). The ``lithium test'' has been successfully applied to identify brown dwarfs in a population of ultracool dwarf stars (Rebolo et al. 1996, Basri 2000). The lithium test is relatively easily applied to M-dwarfs. The reason for this is that the resonance lines of neutral atomic lithium lie in the optical region of the spectrum, at 0.6708 \\mum. One blemish with the ``lithium test'' is the severe blending of lithium lines with the background of TiO lines, nonetheless Li lines break through the molecular background. In the spectra of cooler M and L dwarfs Ti atoms are depleted onto dust particles (Tsuji et al. 1996, Jones \\& Tsuji 1997, Pavlenko 1998) so TiO absorption is weakened. In the L dwarf regime the appearance of lithium lines contends with dust opacity and the wings of K I and Na I resonance lines (see Pavlenko et al. 2000 and references therein). The realisation of the ``deuterium test'' is considerably more difficult. Conventionally, the deuterium abundance of hot objects is determined from analysis of multicomponent features on the background of the $L_\\alpha$ 0.1215 \\mum ~H~I line seen in emission. Unfortunately, this method cannot be used in the case of ultracool dwarfs which are covered by a thick envelope of neutral hydrogen. Observations also yield H~I emission lines ($H_\\alpha$ 0.6563 \\mum) formed in the outermost layers of hot chromospheres or accretion disks around young stars. However, such a plasma is variable and may be polluted by interstellar material. Taking into account these circumstances, it is logical to analyse the spectral lines of deuterated molecules formed in the photospheric layers of ultracool dwarfs. The first investigation and analysis of the combined spectra of \\HHO /HDO were carried out by Chabrier et al. (2000) and Pavlenko (2002). Due to the change in mass and the breakdown of molecular symmetry the vibration-rotation bands of HDO in the mid-infrared spectrum shift with respect to the \\HHO bands. There are a few spectral regions which can be used for detection of HDO lines in the IR spectra of ultracool dwarfs: 3.5 -- 4, $\\sim$5, 6-7 \\mum (Pavlenko 2002). The main problem is that despite the shift in wavelength such HDO lines will be on a background of far stronger \\HHO lines. A possible alternative to HDO are the diatomic hydrides. Strong molecular bands of diatomic metal hydrides such as MgH and CrH can be observed in the optical spectrum of ultracool dwarf stars. The MgH band system \\AMgH - \\XMgH can be observed at 0.47--0.6 \\mum, and the CrH band system \\ACrH -- \\XCrH\\ show absorption features at 0.6--1.5 \\mum. The molecules MgH and CrH have been known in astrophysics for a long time. MgH has been more extensively studied than CrH, because it can be observed in the spectra of G to M type stars. The dissociation energy of MgH is very low (1.285 eV) so lines of this molecule are very sensitive to the temperature and gravity variations in stellar atmospheres. MgH lines were used to determine temperatures in the atmospheres of cool giants (Wyller 1961) and the Sun (Sinha et al. 1979, Sinha \\& Joshi 1982), and for the determination of the surface gravity of stars (Bell \\& Gustaffson 1981, Bell at al. 1985, Berdyugina \\& Savanov 1992, Bonnel \\& Bell 1993). The pure rotational spectrum of MgH and MgD radicals (\\XMgH) in their ground state $v$=0 and $v$=1 vibrational modes has been studied by Ziurys et al. (1993). The first MgH linelist was computed by Kurucz (1993). Recently, more extensive studies of MgH transitions were performed by Weck et al. (2003, 2003a, 2003b) and Skory et al. (2003). Although CrH has been known since Gaydon \\& Pearse (1937), its electronic spectrum remained relatively unstudied for many years. Engvold et al. (1980) identified lines of CrH in a spectrum of sunspots as formed by \\XCrH -- \\XCrH transitions. They used the results of studies of multiplicity of $\\Sigma$-terms of CrH by Klehman \\& Uhler (1959) and O'Connor (1967). Later Ram, Jarman \\& Bernath (1993) performed a rotational analysis of 0--0 band of the \\ACrH -- \\XCrH electronic transition and obtained improved rotational constants for the $v'$ =0 vibrational state. Combining these results with those of Bauschlicher et al. (2001) and Lipus et al. (1991) for the vibrationally excited transitions, Burrows et al.(2002) computed an extended linelist for CrH. Recently Shin et al. (2005) have measured radiative lifetimes of the $v$ = 0,1 levels of \\ACrH state of CrH. These measured lifetimes are about 16\\% -- 45 \\% longer that those obtained by Burrows at al. (2002). These results provide evidence that the oscillator strengths of Burrows et al. (2002) should be corrected by a factor of 0.8 for at least the transitions to the $v'$ = 0. The submillimeter spectra of CrH and CrD formed by pure rotational transitions in the ground electronic state, have been observed in the laboratory by Halfen \\& Ziurys (2004). Electronic bands of MgD and CrD are likely to be located in the same spectral regions as the corresponding bands of MgH and CrH. In this paper we model the bands of these molecules to analyse the possibility of their use for the determination of the D/H ratio in the atmospheres of ultracool dwarfs. In section 2 we present a description of the procedures used to compute the molecular bands of CrH, CrD, MgH and MgD. In section 3 we present the vibrational-rotational constants of MgH, MgD, CrH and CrD. In section 4 we present the results of the computation of molecular bands. In section 5 we discuss the possibility of using the electronic bands of diatomic molecules for a realisation of the ``deuterium test''. ", "conclusions": "The detection of deuterated molecules in the spectra of ultracool dwarfs provide a challenge for both theoreticians and observers. Indeed, in the atmospheres of planetary mass objects (M $<$ 13 M$_J$) we cannot expect ratio D/H $> 2\\times10^{-5}$. This means the lines of deuterated molecules should be about 5000 times weaker than those of the hydrides. The ideal case would be a spectral region where the molecular bands are not blended. So a crucial requirement is a large difference in the wavelengths of the band heads of the hydrides and deuterides. CrH appears to be more useful than MgH in the search for deuterated species. The band heads of CrD are displaced significantly from the band heads of CrH. Bands of CrH are observed in the spectra of the latest L dwarfs (Kirkpatrick et al. 1999). The CrD bands are located in the ``near infrared'' spectra, where fluxes are much higher than in the ``optical'' spectral regions. In this paper we show that the most useful bands for the realisation of the deuterium test are $\\Delta v = +1$ and $\\Delta v = -1$ ($\\lambda$ = 0.795 and 0.968 \\mum, respectively). The $\\Delta v = -1$ band looks especially promising. It is located in the near infrared region with an absence of strong background absorption features. However, portions of this CrD band will be swamped by the 0-0 FeH Wing-Ford band at 0.99$\\mu$m and possibly by water bands. Still, the case for CrD looks better than for HDO lines which are formed on a background of strong H2O lines (Pavlenko 2002). High quality line lists are required to test these possibilities fully. It is worth noting, that a potential problem lies in the possibility that Cr atoms are absorbed onto dust particles. The depletion of Cr will reduce the strength of both CrD and CrH bands. Fortunately as CrH bands are located in the same spectral region, we can ``scale'' the CrD depletion processes by adjusting the strength of CrH bands. For more precise studies, more accurate and detailed linelists of CrH and CrD are required. The calculation of such linelists would require new improved computations supported by new laboratory measurements. One problem that concerns us is that even once we have a good agreement with the model and experimental data for the MgH or CrH molecule, there are still perturbations about which we know very little from experiments done so far. Nonetheless, MgH is a non-starter for the deuterium test and as we note above, the model adopted in our paper is more likely to be reliable for CrH and CrD. It is worth noting that the use of the pure rotational-vibrational bands located in the mid and far infrared spectral region may offer an alternative. Indeed, the displacement between CrH and CrD rotation-vibration bands is even larger, than for the case of electronic bands. Nevertheless, we cannot be certain that we have identified the best candidate systems for the deuterium test. Future investigations of deuterated molecules in different spectral regions are important to determine which offers the best possibility for the realisation of this test. The ideal solution would be to detect lines of deuterated molecule(s) in different spectral regions. This presents a serious observational challenge which can only be met in combination with careful laboratory measurement and the computation of high quality molecular spectra." }, "0710/0710.1472_arXiv.txt": { "abstract": "Very recently, J~1128+5925 was found to show strong intraday variability at radio wavelengths and may be a new source with annual modulation of the timescale of its radio variability. Therefore, its radio variability can be best explained via interstellar scintillation. Here we present the properties of its optical variability for the first time after a monitoring program in 2007 May. Our observations indicate that in this period J~1128+5925 only showed trivial optical variability on internight timescale, and did not show any clear intranight variability. This behavior is quite different from its strong radio intraday variability. Either this object was in a quiescent state in optical in this period, or it is intrinsically not so active in optical as it is in radio regimes. ", "introduction": "Blazars are the most variable subset of AGN. They show a variety of variability timescales. The longest timescales can be far longer than one year, while the shortest may be less than one hour. The variability with a timescale less than one day is often called intraday variability or IDV, as first reported by \\citet{heeschen84}, \\citet{witzel86}, and \\citet{heeschen87}. Strong IDV phenomena have been observed in the radio domain in a large number of blazars. If interpreted as being source intrinsic, the short-timescale variability would require a very small emitting region and hence a very high apparent brightness temperature of $10^{16}\\sim10^{21}\\,\\rm{K}$, which is far beyond the inverse-Compton limit of about $10^{12}\\,\\rm{K}$ \\citep{keller69}. Alternatively, the IDV can be explained via extrinsic origin, e.g., via interstellar scintillation (ISS). A strong support to the ISS origin is the so-called annual modulation of the variability timescale, which is the result of the annual changes of the relative velocity vector between the scattering screen and the Earth as the Earth orbits around the Sun \\citep[e.g.,][]{dennett02,dennett03}. Such annual modulation has been observed in a few IDV sources, as mentioned by \\citet{gabanyi07}. Very recently, the flat-spectrum radio quasar J~1128+5925 was found to show strong IDV at centimeter wavelengths, and its IDV timescale displays an annual modulation \\citep{gabanyi07}. Therefore, its IDV may be caused by ISS. In optical, there is no previous report on its variability. In order to know the properties of its optical variability and to make a comparison to those of its radio variability, we performed an optical monitoring program on this object in 2007 May. Here we present the results. ", "conclusions": "We performed an optical monitoring program on J~1128+5925 in the $R$-band from 2007 May 5 to 29. Our monitoring results indicate that in this period J~1128+5925 only showed trivial optical variability on internight timescale, and did not show any clear intranight variability. Either this object was in a quiescent state in optical regimes in this period, or it is intrinsically not as active at optical as it is at radio wavelengths. Some blazars that show strong IDV in radio regimes also display rapid and strong variability in optical regimes, such as S5~0716+714 \\citep[e.g.,][] {wu05,wu07,montagni06,pollock07}. This doesn't seem to be the case for J~1128+5925. This object exhibits strong IDV at radio wavelengths, but not at optical wavelengths. It is easy to explain this difference if the optical and radio variabilities come from different origins: The optical variability may be intrinsic to the source (the ISS cannot change the optical flux), while the radio variability is mainly the result of ISS, as implied by the observations of \\citet{gabanyi07}. We present the first report on the optical variability of J~1128+5925 in this paper. However, because of the solar conjunction and observations of other targets with the telescope, our monitoring did not last long. More observations are needed to know whether or not this object is always optically quiescent. Multi-band optical monitoring is also necessary to constrain its optical variability in more detail. Of particular interest is to carry out simultaneous optical and radio monitorings on this object in order to make a more direct comparison between the variabilities at these two wavelengths. Future campaigns can investigate whether there is correlated optical IDV when strong radio IDV is observed. If such correlations are detected, it would be strong evidence that both the optical and radio variability structures are intrinsic to the source, as in the case of S5~0716+714 \\citep{quirren91,wagner95,wagner96}. The broadband variability is also helpful to derive for this object some basic parameters, such as the mass of the central supermassive black hole, the boosting factor of the relativistic jet, etc \\citep[e.g.,][]{fan05}." }, "0710/0710.4582_arXiv.txt": { "abstract": "We compare the stellar structure of the isolated, Local Group dwarf galaxy Pegasus (DDO\\,216) with low resolution HI maps from \\cite{young2003}. Our comparison reveals that Pegasus displays the characteristic morphology of ram pressure stripping; in particular, the HI has a ``cometary'' appearance which is not reflected in the regular, elliptical distribution of the stars. This is the first time this phenomenon has been observed in an isolated Local Group galaxy. The density of the medium required to ram pressure strip Pegasus is at least $10^{-5} - 10^{-6}$\\,cm$^{-3}$. We conclude that this is strong evidence for an inter-galactic medium associated with the Local Group. ", "introduction": "\\cite{einasto1974} first highlighted that dwarf satellites of large galaxies tend to be gas deficient compared to isolated dwarfs. The former generally have little or no ongoing star formation and the stars are pressure supported (dwarf spheroidal, dSph). The latter generally have ongoing star formation and the gas dynamics show that rotational support is important (dwarf irregular, dIrr). ``Transition'' dwarfs are gas-rich and, unlike dIrr galaxies, have little or no detectable HII regions, although they usually show indications of recent star formation. The processes by which dwarf galaxies loose their gas are not fully understood. Internal feedback, particularly winds from supernovae, are likely important (\\citealt{dekel1986}) and the existence of the position-morphology relation clearly indicates that environmental influences are significant. \\cite{mayer2006} show that it is possible for dwarf galaxies to be ram pressure stripped of some of their gaseous component in a hot halo of the Milky Way or M31. This idea was originally proposed by \\cite{lin1983}, who calculated the density of the medium required to be of order $10^{-6}$\\,cm$^{-3}$. There have been no direct detections of such a medium, although recently \\cite{nicastro2002,nicastro2003} and \\cite{sembach2003} have detected OVI absorption which they attribute to hot gas associated with either a Milky Way corona or a Local Group medium. In this {\\it Letter}, we compare the stellar and gaseous structure of the isolated, transition-type, dwarf galaxy Pegasus (DDO216). We show that it displays the characteristic signature of ram pressure stripping and conclude that this is strong evidence for hot gas associated with the Local Group. Table~1 summarises some of the observed properties of Pegasus. We adopt the distance estimate by \\cite{mcconnachie2005a}, $D \\simeq 919$\\,kpc, derived from the same photometry used in this Letter. \\begin{table}[htdp] \\begin{center} \\caption{Summary of observed parameters for the Pegasus (DDO216) dwarf galaxy} \\begin{tabular}{lll} Parameter & Value & Reference \\\\ \\hline $\\alpha$ (J2000) & 23h 28m 36.2s& --- \\\\ $\\delta$ (J2000) & +14$^\\circ$ 44$^\\prime$ 35$^{\\prime\\prime}$& --- \\\\ $(l, b)$ & $(94.8^\\circ, -43.6^\\circ)$ & --- \\\\ $M_V~{\\it(L_V)}$ & -12.9~$(1.24 \\times 10^7\\,L_\\odot)$ & \\cite{mateo1998a} \\\\ $M_{HI}$ & $4.06 \\times 10^6\\,M_\\odot$ & \\cite{young2003}\\\\ $v_\\odot$ & -183\\,km\\,s$^{-1}$ & \\cite{young2003} \\\\ $v_r/\\sigma$ & $1.7$ & \\cite{mateo1998a} \\\\ Distance & 24.82 $\\pm$ 0.07 (919\\,kpc) & \\cite{mcconnachie2005a} \\\\ & 24.4 $\\pm$ 0.2 & \\cite{gallagher1998} \\\\ & 24.9 $\\pm$ 0.1 & \\cite{aparicio1994} \\\\ \\hline \\end{tabular} \\end{center} \\label{distances} \\end{table} ", "conclusions": "\\subsection{Comparison of stellar and HI contours} The top-right panel of Figure~1 shows the tangent-plane projection of the spatial distribution of objects identified as stellar from our INT~WFC observations of Pegasus. The dotted lines in this panel (and the remaining panels of Figure~1) correspond to the approximate edges of each CCD of the INT~WFC. Only objects which lie within $1-\\sigma$ of the stellar locus in both the $V^\\prime$ and $i^\\prime-$band observations are shown. The hole at the center of the main body of Pegasus is due to severe crowding which causes incompleteness. The bottom-left panel of Figure~1 shows a contour map of the density distribution of stars. The first contour is $2-\\sigma$ above the background, and the contours correspond to $2.2, 5.0, 8.6, 13.2, 19,0, 26.3, 35.7, 47.5$ and $62.5$\\,stars\\,arcmin$^{-2}$. The contour map was made in the standard way and follows exactly the methodology described in \\cite{mcconnachie2006b}. This panel shows that Pegasus is significantly more extended than suggested by the image in the first panel. The bottom-right panel of Figure~1 shows the stellar density distribution as a grey-scale with square-root scaling. The red contours are the low-resolution HI distribution from \\cite{young2003}. The contours correspond to column densities of $0.1, 0.2, 0.4, 0.8, 1.6, 3.2, 6.4$ and $12.8 \\times 10^{20}$\\,cm$^{-2}$. Whereas the stars are distributed in a regular ellipse (typical of a flattened spheroid or an inclined disk) the HI has a ``cometary'' appearance; the contours in the south-east are more closely packed and do not extend as far as in the north-west. \\subsection{Is Pegasus being ram pressure stripped?} \\begin{figure*} \\begin{center} \\includegraphics[angle=270, width=8.9cm]{f1a.eps} \\includegraphics[angle=270, width=8.9cm]{f1b.ps} \\includegraphics[angle=270, width=8.9cm]{f1c.ps} \\includegraphics[angle=270, width=8.9cm]{f1d.ps} \\caption{Projections in the tangent plane $\\left(\\xi, \\eta\\right)$ of the structure of the Pegasus dwarf galaxy (DDO216) with the orientation of the field indicated. The dotted lines trace the approximate edges of the four CCDs of the INT~WFC. Top-left panel: The reduced $V^\\prime-$band image of Pegasus taken with the INT~WFC. Also marked are the positions of fields analysed in previous studies of this galaxy; the fields analysed by \\cite{gallagher1998} are shown in green as the largest rectangular field and the small WFPC2 footprint, the field analysed by \\cite{aparicio1994} is shown in blue as the smallest rectangular field, and the field studied by \\cite{hoessel1982} is shown in red as the medium sized rectangle. Top-right panel: The distribution of all sources confidently identified as stellar in both the $V^\\prime$ and $i^\\prime-$bands. The hole at the center of Pegasus is due to severe crowding causing the photometry to become seriously incomplete. Bottom-left panel: The stellar density distribution of Pegasus shown as a contour map. The first contour is $2-\\sigma$ above the background and the 9 contours correspond to $2.2, 5.0, 8.6, 13.2, 19,0, 26.3, 35.7, 47.5$ and $62.5$\\,stars\\,arcmin$^{-2}$. Bottom-right panel: the stellar density distribution is shown as a grey-scale with square-root scaling. The red contours show the low resultion HI distribution from \\cite{young2003}. Contours correspond to $0.1, 0.2, 0.4, 0.8, 1.6, 3.2, 6.4$ and $12.8 \\times 10^{20}$\\,cm$^{-2}$. Also shown are the projected directions to all galaxies within $\\sim 500$\\,kpc of Pegasus which have a significant gaseous content. Pegasus displays the characteristic morphology of ram pressure stripping.} \\label{fig} \\end{center} \\end{figure*} The shape of the low resolution HI contours in Pegasus - the smooth, compressed contours in the south-east and the ``tail'' to the north-west - is very similar to the simulated morphology of gas undergoing ram pressure striping (e.g., \\citealt{stevens1999,mori2000,marcolini2003,roediger2005,mayer2006}). Observationally, the M81 group dwarf galaxy Holmberg~II is observed to have a similar morphology (\\citealt{bureau2002}), interpreted as evidence of an intra-group medium. In clusters of galaxies, ram pressure stripping of galaxies by an intra-cluster medium is used to explain various observations, including the deficit of HI in cluster spiral galaxies compared to field spirals (e.g., \\citealt{giovanelli1985}). Indeed, several individual galaxies in the Virgo Cluster have been shown to display gaseous morphologies indicative of ram-pressure stripping (\\citealt{vollmer2000,vollmer2004,vollmer2005}). What else could explain the peculiar appearance of Pegasus? While tidal stripping by large galaxies can affect the structure of dwarf galaxies (e.g., \\citealt{penarrubia2007b}), the closest large galaxy to Pegasus is M31 at $\\sim 470$\\,kpc (all the distance estimates in Table~1 place Pegasus at $> 400\\,$kpc from M31). Even if we assume Pegasus is a weakly-bound satellite of M31, tidal effects at this distance are minimal. If Pegasus was disrupted at pericenter, it is unlikely that the gas would still show signs of current disturbance. Further, tidal stripping tends to produce symmetrical distortions and both gas and stars should be affected. However, these are inconsistent with the structure of Pegasus that we observe. Could the appearance of Pegasus be due to internal effects rather than external influences? Enhanced star formation in the south-east of Pegasus could produce winds which remove gas from this region. However, if this is the case then the densely packed contours in the south-east should have a more concave, rather than convex, shape. For example, \\cite{young2007} discuss a gas cloud associated with the Phoenix dwarf galaxy and conclude that it was blown out by supernovae winds based in part on the concave shape of its contours. An alternative explanation for the HI morphology of Pegasus is that it consists of multiple HI clouds, the sum total of which has a cometary appearance. Figure~6 of \\cite{young2003} is a position-velocity diagram of Pegasus along its major axis. It shows a gradient in velocity and two main concentrations of HI which \\cite{young2003} interpret as two distinct HI clouds. The strength of the secondary feature ($v \\sim -200$\\,km\\,s$^{-1}$) is weaker than the main feature ($v \\sim -180$\\,km\\,s$^{-1}$) and they join at relatively high column density (between the $8 - 16 - \\sigma$ contour levels). An alternative explanation of the data is that the overall velocity gradient is a result of ram-pressure stripping. The velocity difference between the two features may be due to stripped gas leaving a ``hole'' in the distribution, making the secondary feature appear at a higher density than its immediate surroundings (we do not necessarily expect that the column density should smoothly vary over the entire cloud). Henceforth, we adopt the hypothesis that Pegasus is being ram pressure stripped. Following \\cite{gunn1972}, material will be ram pressure stripped from a galaxy if the density of the surrounding medium, $n_{IGM} \\gtrsim \\left(2\\,\\pi\\,G\\,\\Sigma_{T}\\,\\Sigma_{HI}\\right)/\\left(\\mu v^2\\right)$. $\\Sigma_{T}$ is the total surface density (stars plus gas), $\\Sigma_{HI}$ is the column density of HI and $v$ is the relative velocity of the galaxy to the medium. Thus, \\begin{eqnarray} n_{IGM} &\\simeq& 3.7 \\times 10^{-6}\\,\\rm{cm}^{-3} \\left(\\frac{\\rm{100\\,km\\,s^{-1}}}{v}\\right)^2 \\nonumber\\\\ &&\\left(\\frac{\\Sigma_{HI}}{0.1 \\times 10^{20}\\,\\rm{cm}^{-2}}\\right)^2~, \\end{eqnarray} \\noindent where we take the mean particle mass $\\mu = 0.75\\,m_p$ for fully ionized media. We approximate the Local Group space velocity of Pegasus as $v \\sim \\sqrt{3}\\,\\sigma_{LG} \\sim 100$\\,km\\,s$^{-1}$ where $\\sigma_{LG} \\sim 60$\\,km\\,s$^{-1}$ is the Local Group line-of-sight velocity dispersion (\\citealt{sandage1986}). HI at a column density much lower than $\\Sigma_{HI} \\sim 0.1 \\times 10^{20}$\\,cm$^{-2}$ has been stripped from Pegasus, implying that this is a reasonable lower limit for use in this calculation. We adopt $\\Sigma_T = \\Sigma_{HI} \\left(1 + M_\\star/M_{HI}\\right)$, where $M_\\star \\sim 1.24 \\times 10^7\\,M_\\odot$ is the stellar mass of Pegasus (Table~1). This seems reasonable; the surface brightness of Pegasus is $25$\\,mags\\,arcsec$^{-2}$ at a radius of $r = 1.5^\\prime$ on the minor axis (\\citealt{nilson1973,devaucouleurs1991}), corresponding to a stellar surface density of $\\Sigma_\\star \\sim 4 \\times 10^{20}$\\,cm$^{-2}$. This is approximately equivalent to the stellar-to-gas mass ratio multiplied by the gas surface density ($M_\\star/M_{HI} \\times \\Sigma_{HI}$) at $r = 1.5^\\prime$. These values yield $n_{IGM} \\sim 3.7 \\times 10^{-6}$\\,cm$^{-3}$. However, given the uncertainties involved, it is entirely plausible that the value of $n_{IGM}$ could be at least an order of magnitude larger than in Equation~1. \\subsection{Consequences} What is the source of the material that is stripping Pegasus? The bottom right panel of Figure~1 shows the distances of Pegasus to its nearest gas-rich neighbours. The dwarf neighbours are unlikely to be the source of the stripping medium; not only is the required mass of gas unrealistically large (an ejected spherical shell $\\sim 1$\\,kpc thick with a radius of $\\sim 300$\\,kpc would have a mass $> 3 \\times 10^6\\,M_\\odot$ at a density of $n_{IGM}$) but the energy required is too large for a dwarf galaxy to reasonably provide. Alternatively, the gas could be associated with M31. From observations of the Magellanic stream, \\cite{murali2000} estimate that the density of the Milky Way halo at the stream must be $\\lesssim 10^{-5}$cm$^{-3}$, although \\cite{stanimirovic2002} estimate $\\sim 10^{-4}$ cm$^{-3}$. If the gas density in the halo of M31 is similar, then not only would M31 need to have a very extended corona, but its density would need to decrease very slowly with radius. Indeed, if the Milky Way has a similarly extended corona, then the two will overlap and the result may be observationally indistinguishable from a Local Group medium. The isolation of Pegasus raises the strong possibility that the stripping medium is associated with the Local Group, rather than individual galaxies within the group. Clusters of galaxies have such media, and observations of Holmberg~II imply the presence of an intra-group medium in the M81 group (\\citealt{bureau2002}). The density of the intra-group medium implied in Equation~1 is of the same order as the density of the medium responsible for local OVI absorption detected by \\cite{nicastro2002,nicastro2003} and \\cite{sembach2003}, which they suggest is associated with either a Milky Way corona or a Local Group medium. Our result favors the latter interpretation. Theoretically, $\\sim 30\\,\\%$ of baryons in the Local Volume are expected to be in a warm/hot phase ($T \\sim 10^5 - 10^6\\,$K; \\citealt{kravtsov2002}); this is likely concentrated around galaxies and galaxy groups as an intra-group medium. If the stripping medium pervades the Local Group, why do more dwarf galaxies not show evidence of ram pressure stripping? \\cite{lin1983} suggest that all the dSphs have been stripped in this fashion, (although \\cite{mayer2006} show that ram-pressure stripping by itself is insufficient to remove all the gas from a dIrr). It is possible that the Local Group medium will be clumpy and perhaps Pegasus is passing through a region of higher density compared to other dIrrs. Alternatively, Pegasus could be falling into and interacting with the Local Group for the first time, as has recently been speculated for two dSph galaxies at large radii from M31 (And~XII, \\citealt{chapman2007}; AndXIV, \\citealt{majewski2007}). However, the reason why only Pegasus currently shows signs of ram-pressure stripping is unlikely to be known until such time as the masses and orbits of the dIrrs have been determined. Given the distances of these galaxies, this will be some time yet." }, "0710/0710.3477_arXiv.txt": { "abstract": "We use Monte-Carlo simulations, combined with homogeneously determined age and mass distributions based on multi-wavelength photometry, to constrain the cluster formation history and the rate of bound cluster disruption in the Large Magellanic Cloud (LMC) star cluster system. We evolve synthetic star cluster systems formed with a power-law initial cluster mass function (ICMF) of spectral index $\\alpha =-2$ assuming different {\\dts}s. For each of these cluster disruption time-scales we derive the corresponding cluster formation rate (CFR) required to reproduce the observed cluster age distribution. We then compare, in a ``Poissonian'' $\\chi^2$ sense, model mass distributions and model two-dimensional distributions in log(mass) vs. log(age) space of the detected surviving clusters to the observations. Because of the bright detection limit ($M_V^{\\rm lim} \\simeq -4.7$ mag) above which the observed cluster sample is complete, one cannot constrain the characteristic cluster disruption time-scale for a $10^4$ M$_\\odot$ cluster, $t_4^{\\rm dis}$ (where the disruption time-scale depends on cluster mass as $t_{\\rm dis} = t_4^{\\rm dis} (M_{\\rm cl} / 10^4 {\\rm M}_\\odot )^\\gamma$, with $\\gamma \\simeq 0.62$), to better than a lower limit, $t_4^{\\rm dis} \\ge 1$\\,Gyr. \\\\ We conclude that the CFR has been increasing steadily from 0.3 clusters Myr$^{-1}$ 5 Gyr ago, to a present rate of $(20-30)$ clusters Myr$^{-1}$, for clusters spanning a mass range of $\\sim 100-10^7$ M$_\\odot$. For older ages the derived CFR depends sensitively on our assumption of the underlying CMF shape. If we assume a universal Gaussian ICMF, then the CFR has increased steadily over a Hubble time from $\\sim 1$ cluster Gyr$^{-1}$ 15 Gyr ago to its present value. On the other hand, if the ICMF has always been a power law with a slope close to $\\alpha=-2$, the CFR exhibits a minimum some 5 Gyr ago, which we tentatively identify with the well-known age gap in the LMC's cluster age distribution. ", "introduction": "\\label{sec:intro} The mass and age distributions of star cluster systems contain the (fossil) records of their formation conditions. They are therefore among the best tracers of the star-formation histories of their host galaxies available to observers. It is important to realise, however, that one needs to understand both the dominant internal and external evolutionary processes in order to disentangle this formation record, and hence obtain a glimpse of the initial conditions required for star cluster formation. The effects of stellar evolution in a given star cluster (which can be approximated by a ``simple'' stellar population once the cluster has reached an age that is well in excess of its formation time-scale) are rather well understood, whereas we have only recently begun to make major {\\it quantitative} inroads into understanding the environmental effects leading to star cluster ``weight loss'' (i.e., the preferential depletion of the low-mass component of the cluster's stellar mass function caused by tidal stripping and the ejection of stars owing to internal two-body relaxation) and -- eventually -- disruption. Estimates of the characteristic cluster disruption time-scales in various star cluster environments have been calculated by Boutloukos \\& Lamers (2003), de Grijs et al. (2003a,b,c), Gieles et al. (2005) and de Grijs \\& Anders (2006), among others (see also Lamers et al. 2005a,b). Specifically, Boutloukos \\& Lamers (2003) and Lamers et al. (2005b) show that the cluster age distribution and cluster mass function approximate a double power law when a cluster system is affected by both fading and secular dynamical evolution. Knowledge of the detection limit in terms of cluster mass vs. age, combined with the estimate of the cluster age or mass at the ``break point'' of the integrated cluster age (mass) distribution (see, e.g., fig. 1 in Boutloukos \\& Lamers 2003, where the ``break points'' are referred to as $t_{\\rm cross}$ and $M_{\\rm cross}$ in the age and mass distributions, respectively) then leads to the typical cluster disruption time-scale. Their analysis, however, explicitly builds on the assumption of a constant cluster formation rate (CFR, i.e. the number of clusters formed per linear time interval ${\\rm d}N/{\\rm d}t$ is constant in time) as a function of time. In that context, the (poorly-known) time-variable CFR of the LMC cluster system hampers such an analysis. If only the observed cluster age distribution is known, then we are left with a degeneracy between the CFR and the disruption time-scale, in the sense that one cannot distinguish between a low CFR combined with slow secular dynamical evolution on the one hand, and a vigourous CFR combined with cluster disruption occurring on a rapid time-scale on the other. The star cluster system in the Large Magellanic Cloud (LMC) has the potential of providing strong constraints to the theory of star cluster disruption as a function of environment, since it is composed of the largest resolved cluster system spanning {\\it both} a reasonable mass range ($\\sim 10^2 - 10^6$ M$_\\odot$; cf. Hunter et al. 2003, hereafter H03; de Grijs \\& Anders 2006, and references therein) {\\it and} an age range from a few Myr to $\\sim 13$ Gyr available. In addition, thanks to the LMC's proximity, we have been able to obtain observations -- and derived the age and mass distributions -- of a sufficiently large cluster sample to allow a statistical approach to its evolution (e.g., H03; de Grijs \\& Anders 2006; see also Sect. 2). On the basis of the few star clusters systems analysed in detail to date, including M51 and the Antennae interacting system, Bastian et al. (2005), Fall et al. (2005) and Fall (2006) suggest that the early evolution of star cluster systems is most likely characterised by a rapid, largely mass-independent ``infant mortality'' phase, at least for masses $\\ga 10^4$ M$_\\odot$ (see also de Grijs \\& Parmentier 2007; and references therein), combined with ``infant weight loss''(the loss of stars caused by rapid, early gas expulsion; cf. Weidner et al. 2007), the effects of which are enhanced by stellar evolutionary mass loss. In this scenario, this early phase, which ends when clusters reach an age of $\\sim 40$ to 50 Myr (e.g., Goodwin \\& Bastian 2006), would then be followed by (mass-dependent) secular evolution. The early, rapid cluster disruption process results from the expulsion of the intracluster gas due to adiabatic or explosive expansion driven by stellar winds or supernova activity (Mengel et al. 2005; Bastian \\& Goodwin 2006; Goodwin \\& Bastian 2006; see de Grijs \\& Parmentier 2007 for a review). Star clusters are expected to settle back into virial equilibrium $\\sim 40$ to 50 Myr after gas expulsion (Goodwin \\& Bastian 2006). In our analysis in this paper we will therefore exclude star clusters younger than 50 Myr, since our main purpose is to derive the characteristic (mass-dependent) time-scale of cluster disruption in the LMC driven by secular evolution only. In a follow-up paper (Goodwin et al., in prep.), we will discuss the evolution of the LMC cluster system on the shortest time-scales relevant to the infant mortality and infant weight loss scenarios. In de Grijs \\& Anders (2006), we found that the LMC's CFR has been roughly constant outside of the well-known age gap between $\\sim 3$ and 13 Gyr, when the CFR was a factor of $\\sim 5$ lower (assuming a roughly constant rate during this entire period). Based on this observation as our main underlying assumption, we used a simple approach to derive the characteristic cluster disruption time-scale in the LMC, for which we found that $\\log(t_4^{\\rm dis} {\\rm yr}^{-1}) = 9.9 \\pm 0.1$, where $t_{\\rm dis} = t_4^{\\rm dis} (M_{\\rm cl}/10^4 {\\rm M}_\\odot)^{0.62}$ (Boutloukos \\& Lamers 2003; Baumgardt \\& Makino 2003; Gieles et al.~2005). We argued that this was consistent with earlier, preliminary work done on a smaller cluster sample: Boutloukos \\& Lamers (2002) found $\\log( t_4^{\\rm dis} {\\rm yr}^{-1} ) = 9.7 \\pm 0.3$ for a smaller sample of 478 clusters within 5 kpc from the centre of the LMC, in the age range $7.8 \\le \\log(\\mbox{age yr}^{-1}) \\le 10.0$. We also considered our result qualitatively consistent with Hunter et al. (2003), who noticed very little destruction of clusters at the high-mass end. This long characteristic disruption time-scale would imply that hardly any of our LMC sample clusters are affected by significant disruptive processes, so that we are in fact observing the {\\it initial} cluster mass function (CMF). However, a close inspection of fig. 6 of de Grijs \\& Anders (2006) highlights an apparent contradiction. The ``crossing time'', $t_{\\rm cross}$, defined by the crossing point between the best-fitting lines describing the number of clusters per unit time-scale that are mostly affected by fading of their stellar populations and those that are undergoing significant secular disruption, seems to imply that a more appropriate time-scale for the disruption of the LMC cluster system may be of order $\\log( t_4^{\\rm dis} {\\rm yr}^{-1} ) \\simeq 8.9$. Since this implies a downward adjustment of the characteristic cluster disruption time-scale in the LMC by up to an order of magnitude, we decided to re-investigate the LMC's cluster disruption history. Here, we approach this problem from a different angle, by running a large number of Monte-Carlo simulations in which we vary the cluster disruption time-scale. Meanwhile, for each of these cluster disruption time-scales we derive the corresponding CFR required to reproduce the observed cluster age distribution. We then match the observed cluster mass distribution, integrated over time, and the observed two-dimensional distribution of the detected surviving clusters in the log(mass) vs. log(age) plane to the model results. $\\chi^2$ fit estimates are used to quantify which cluster disruption time-scale and, therefore, which cluster formation history, best describes the presently available data. This paper is organised as follows. In Section \\ref{sec:data} we justify our choices used in the data analysis leading to the cluster age and mass distributions used in the remainder of the paper. Section \\ref{sec:synth} discusses our basic assumptions in constructing synthetic cluster populations, which we then use in Section \\ref{sec:disr} to explore the range of characteristic cluster disruption time-scales allowed by the data. In Section \\ref{sec:comp} we highlight the importance of properly understanding the data's completeness limit, and use this in Section \\ref{sec:agap} to constrain possible variations in the CFR over time. In Section \\ref{sec:1vs10}, we assess precisely what we would need to fully and unambiguously constrain $t_4^{\\rm dis}$. Our results and conclusions are summarized in Section \\ref{sec:conc}. \\begin{figure*} \\begin{minipage}[t]{\\linewidth} \\centering\\epsfig{figure=obs_age_mass.eps, width=\\linewidth} \\end{minipage} \\caption{Distribution of the LMC star cluster sample of de Grijs \\& Anders (2006) in the [$\\log({\\rm age}),\\log(M_{\\rm cl})$] plane. The filled triangles correspond to the vertical dashed lines in the individual panels of Fig.~\\ref{fig:compl} (upright triangles: left-hand column; upside-down triangles: right-hand column). For subsequent cluster age ranges (in steps of 0.25 dex and 0.5 dex wide) they trace the mass limit below which the sample becomes incomplete (see section \\ref{sec:synth} for details). They are therefore considered tracers of the fiducial detection limit (thick dash-dotted line with squares), which corresponds to $M_V^{\\rm lim}=-4.7$ mag (based on the {\\sc galev} mass-to-light ratios for ``simple'' stellar populations). The three thin dash-dotted lines, labelled `[1]', `[2]' and `[3]', are the detection limits corresponding to $M_V^{\\rm lim}=-5.2$, $M_V^{\\rm lim}=-4.2$ and $M_V^{\\rm lim}=-3.5$ mag, respectively. They are therefore equivalent to the thick dash-dotted line shifted vertically by, respectively, $\\Delta \\log (M_{\\rm cl})=0.2, -0.2$, and $-0.48$. The lower dash-dotted curve (`[3]') is the $M_V^{\\rm lim}=-3.5$\\,mag fading limit of H03. The thick dashed lines represent the cluster disruption limits for $\\log (t_4^{\\rm dis} {\\rm yr}^{-1})=8.1, 9.0$ and $9.9$ (labelled `[a]', `[b]' and `[c]', respectively). The age range on which we focus in this paper is bracketed by vertical solid lines, at $\\log (\\mbox{age yr}^{-1})=7.7$ and $\\log (\\mbox{age yr}^{-1})=9.2$; clusters brighter than $M_V^{\\rm lim}=-4.7$ mag in that age range are represented by crosses} \\label{fig:obs_am} \\end{figure*} \\begin{figure*} \\begin{minipage}[t]{\\linewidth} \\centering\\epsfig{figure=MF_agebin_completeness.eps, width=\\linewidth} \\end{minipage} \\caption{Observed cluster mass functions for the age ranges included at the top of each panel. In each panel, the vertical dashed line is the mass limit bracketing 25 and 75 per cent of the cluster subsample on either side. This is a good proxy to the cluster mass at the turn-over of each CMF. For each age range, turn-over masses are indicated in ($\\log (\\rm age),\\log (M_{\\rm cl})$) space as filled triangles in Fig.~\\ref{fig:obs_am}. The mass limits defined by the dashed lines evolve with time following a line of constant luminosity, at $M_V=-4.7$ mag. This implies that the decrease in cluster numbers observed for each CMF below its turn-over mass (i.e. below the vertical dashed line) is mostly driven by incompleteness effects} \\label{fig:compl} \\end{figure*} ", "conclusions": "\\label{sec:conc} In this paper, we have carried out numerous detailed Monte-Carlo simulations aimed at constraining the cluster formation history and the rate of bound cluster disruption in the LMC star cluster system. We considered only clusters older than 50\\,Myr in order not to determine erroneously short cluster disruption time-scale as a result of the inclusion of the infant mortality and infant weight loss evolutionary phases. Our data are based on the $UBVR$ photometry of H03 (obtained from Massey's (2002) survey of the Magellanic Clouds), for which de Grijs \\& Anders (2006) obtained homogeneously determined age and mass estimates. We estimated a fiducial detection limit above which the cluster sample is (fairly) complete from the CMF as a function of age, at $M_V^{\\rm lim} = -4.7 {\\rm mag}$. This is significantly brighter than H03's fading limit. We consider only clusters brighter than this limit, in order to avoid severe statistical incompleteness effects. We have evolved synthetic star cluster systems characterised by constant CFRs assuming 20 different {\\dts}s. The CFR was adjusted to reproduce the observed age distribution. By doing so, we are likely to loose sensitivity to CFR variations occurring on a time-scale shorter than the width of the age range corresponding to each logarithmic age bin (i.e., more likely at old age where our age bins span greater linear age ranges than at young age). However, the general behaviour of the CFR, as well as the CFR averaged over age are robustly recovered (see Section \\ref{sec:1vs10}). We then compare, in a ``Poissonian'' $\\chi^2$ sense the modelled mass distribution and the modelled [$\\log({\\rm age}),\\log(M_{\\rm cl})$] distribution to the observations. We show that because of the bright detection limit at $M_V^{\\rm lim}= -4.7$ mag, one cannot constrain $t_4^{\\rm dis}$ to better than a lower limit, $t_4^{\\rm dis} \\ge 1$\\,Gyr. The tightest constraints are set by the CMF integrated over age. The $\\chi^2$ test applied to the distribution of points in [$\\log({\\rm age}),\\log(M_{\\rm cl})$] space turns out to be a poor diagnostic tool. This is probably related to the low density of points in each cell of the [$\\log({\\rm age}),\\log(M_{\\rm cl})$], compared to the density of points in each bin of the integrated CMF, leading to smaller Poissonnian error bars and therefore a better constrained model in the latter case. Our range of {\\dts} estimates is robust with respect to model variations, such as of the upper limit of the initial cluster mass range, the location of the grid in [$\\log({\\rm age}),\\log(M_{\\rm cl})$], and the size of cells in this grid. We have shown that should the detection limit be underestimated, artificially shortened {\\dts}s would result. This is so because there is a degeneracy between incompleteness and secular evolution, i.e., the fading-driven turn-over in a CMF in a given age bin is interpreted as resulting from secular evolution, leading to a shortened cluster disruption time-scale. Having set the best possible constraints on $t_4^{\\rm dis}$, we explored the corresponding CFR, in particular considering $t_4^{\\rm dis}=1$ and 10 Gyr. The CFR has been increasing steadily from about 0.3 clusters Myr$^{-1}$ 5 Gyr ago, to a present rate of $(20-30)$ clusters Myr$^{-1}$, for clusters spanning an initial mass range of $\\sim 100-10^7$ M$_\\odot$. The CFRs inferred for both disruption time-scales differ by at most a factor of three (Fig.~\\ref{fig:ad_all}). At older ages however, the situation becomes unclear. The uncertainty in the CFR as a result of the uncertainty on $t_4^{\\rm dis}$ increases. In addition, the overall temporal behaviour of the CFR depends on the shape of the ICMF of the oldest, globular cluster-like objects. If this is the universal Gaussian ICMF, then the CFR has increased steadily over a Hubble time from $\\sim 1$ cluster Gyr$^{-1}$ 15 Gyr ago to its present value. On the contrary, if the ICMF has always been a power law with a slope close to $\\alpha=-2$, the CFR exhibits a minimum some 5 Gyr ago. Our results may be related to the orbital history and dynamics of the LMC with respect to both the SMC and the Milky Way, although this remains poorly constrained because of a lack of proper motion measurements with the required accuracy (Besla et al.~2007). Additionally, we note that interactions between the Clouds and between the Clouds and the Galaxy, while affecting their star-formation history, also affect the {\\it structure} of the Magellanic Clouds (Bekki \\& Chiba 2005). This may have induced temporal variations in the {\\dts} over the past Hubble time, rendering the {\\dts} estimate at old age even more uncertain. Finally, we have investigated which strategy should be adopted in the future in order to better constrain the {\\dts} in the LMC. Specifically, we have generated synthetic cluster populations defined by a given cluster formation history, ICMF and various dissolution rates. After the addition of Gaussian noise to mimic an observational situation, we processed these simulated data sets in the same way as the actual LMC data. We demonstrate that {\\it if} the {\\dts} is known, then the CFR can be derived accurately. We confirm our inability to distinguish $t_4^{\\rm dis}=1$ Gyr from 10 Gyr because of the bright detection limit. With such a bright detection limit, the expected turn-over in the CMF caused by dynamical evolution is not detected, for any cluster age range. As a result, only a lower limit to the {\\dts} can be retrieved, i.e., we can exclude all $t_4^{\\rm dis}$ that are short enough to lead to a turn-over above the detection limit. To obtain age and mass estimates for an LMC star cluster sample complete above $M_V^{\\rm lim}=-3.5$ is desirable to more easily distinguish between $t_4^{\\rm dis}=1$ Gyr and $t_4^{\\rm dis}=10$ Gyr." }, "0710/0710.3180.txt": { "abstract": "We present new optical spectroscopy and photometry, 2MASS infrared observations and 24 years of combined AAVSO and AFOEV photometry of the symbiotic star candidate \\ae. The long-term light curve is characterized by outbursts lasting several years and having a slow decline of $\\sim 2 \\times 10^{-4}$ mag/day. The whole range of variability of the star in the $V$ band is about 4 magnitudes. The periodogram of the photometric data reveals strong signals at $\\sim$ 342 and 171 days. The presence of the emission feature at $\\lambda$ 6830 \\AA~ at minimum and the detection of absorption lines of a $\\sim$ K5 type star confirm the symbiotic classification and suggest that AE\\,Cir is a new member of the small group of s-type yellow symbiotic stars. We estimate a distance of 9.4 kpc. Our spectrum taken at the high state shows a much flatter spectral energy distribution, the disappearance of the $\\lambda$ 6830 \\AA~ emission feature and the weakness of the He\\,II 4686 emission relative to the Balmer emission lines. Our observations indicate the presence of emission line flickering in time scales of minutes in 2001. The peculiar character of \\ae~ is revealed in the visibility of the secondary star at the high and low state, the light curve resembling a dwarf nova superoutburst and the relatively short low states. The data are hard to reconciliate with standard models for symbiotic star outbursts. %Another peculiarity is the detection of features of the cool star at the low and high states. %We discuss the multi-wavelength data and suggest that the ellipsoidal-like variability is due to %fluorescence of the cool star atmosphere that is irradiated by high energy photons from the hot component. %We observe a 40-day length eclipse-like episode in %our photometry, that we interpret as self-occultation of the irradiated hemisphere of the red star. Then, assuming a white dwarf %hot component of 1 M$\\odot$ we find that the red giant probably fills their Roche lobe and has a radius $R_{g} \\approx$ 100 R$_{\\odot}$. %From independent methods we have found a distance to AE\\,Cir of 14 $\\pm$ 3 kpc. %Our observations suggest that accretion onto a compact object %in a semi-detached binary could play a major role in the high states and short-term variability observed in AE\\,Cir. ", "introduction": "AE\\,Cir (S32 = StHA32, $\\alpha_{2000}$ = 14:44:52.0, $\\delta_{2000}$ = -69:23:35.9) was classified as a RCB star in the Fourth General Catalog of Variable Stars. This classification was rejected by Kilkenny (1989), who found strong emission lines of H\\,I and weaker emission lines of He\\,II and other species in one spectrogram spanning a wavelength range of 3400-5100 \\AA. Kilkenny noted the lack of forbidden lines and the anomalous strong He\\,II 4686 emission for this object, and suggested a symbiotic classification, although no lines of the cool component were detected. He also mentions the photometric variability of the star, ranging from 12 to 14 mag. (from visual estimates) in hundreds of days. $B,V$ photoelectric measures by Lawson \\& Cottrell (1990) taken in an interval of 107 days show the star with $V$ magnitude between 13.52 and 14.51 and color $B-V$ between 0.95 and 1.43, the object being redder when fainter. Based on the spectroscopy by Kilkenny, AE\\,Cir was listed as a suspected symbiotic in the catalog of Belczynski et al. (2000). Symbiotic stars have been reviewed recently by Miko{\\l}ajewska (2007). In this paper we present new photometry and low and high-resolution spectra of AE\\,Cir and investigate their emission line properties and symbiotic nature. In Section 2 we give details of our spectroscopic observations and show the methods used in data reduction and analysis. In that section we also analyze available visual long-term photometric records along with our own CCD photometry. In Section 3 we analyze our data, giving the main results of our research. In Section 4 we discuss our results in the context of symbiotic stars and present our conclusions in Section 5. ", "conclusions": "In this paper we have investigated the nature of the symbiotic star candidate AE\\,Cir. We have analyzed new optical photometric and spectroscopic data, 2MASS infrared photometry and 24 years of visual photometry in an integrated way in order to get a better understanding of the system. Our conclusions can be summarized as follows:\\\\ \\begin{itemize} \\item The symbiotic nature is confirmed. This result is based on the detection of the $\\lambda$ 6830 \\AA~ emission feature and the spectral signatures of a cool stellar component of spectral type $\\sim$ K5. \\item The spectral type $\\sim$ K5 and the spectral and photometric properties indicate that \\ae~ is a member of the small group of yellow symbiotic stars of the s-type. \\item The light curve is characterized by outbursts lasting $\\sim$ 4000 days and overall amplitude of variability about 4 magnitudes. The outbursts show a slow decline of $\\sim 2 \\times 10^{-4}$ mag/day. The duration of the low state is about 38\\% the high state. \\item A strong signal at 342 $\\pm$ 15 days is detected in the light curve. The light curve folded with this period shows two broad minima with different amplitude. \\item The spectrum in the low state is characteristic of a symbiotic star without forbidden lines and very strong He\\,I\\,4868 emission. At maximum the emission feature at $\\lambda$ 6380 \\AA~ disappears, the spectrum becomes flat and the relative intensity of the He\\,I 4686 \\AA~line, compared with the Balmer lines, becomes lower. At the same time the overall emission line spectrum shows smaller equivalent widths. Signatures of the cool companion are observed at the high and low state. \\item We observe H\\,I, He\\,I and He\\,II emission line variability in time scales of minutes in 2001 being larger in the line centers. The Na\\,D and $\\lambda$ 6820 \\AA~ lines do not follow this variability. We also observe larger broadening and asymmetry in the Balmer lines compared with others emission lines in this epoch. \\item The data suggest that full visibility of the red giant should occur at the low state at $V \\sim 15.5$. For a typical K5 giant behind the galactic disc this would indicate a distance of 9.4 kpc. This should indicate that \\ae~ is about 1.4 kpc below the galactic plane. \\item We suggest that an eclipse-like event could be interpreted as self-occultation of the irradiated red giant hemisphere. In this case we found that it is possible that the red giant fills its Roche-lobe. Assuming a white dwarf hot component of 1 M$_{\\odot}$, we estimate a radius of 99 R$_{\\odot}$, a mass of 1.1 M$_{\\odot}$ and M$_{V}$ = -2.4 for the red giant at the epoch of the SMARTS observations. \\item The atypical character of \\ae~ is revealed in the long time passing in outburst and the dwarf nova superoutburst shape of the light curve. Another atypical feature is the rather strong TiO absorption band around $\\lambda$ 6300 \\AA~ at the high state. This could indicate that the red giant and the hot component are partially obscured at minimum or the brightness of the secondary star is not constant. Obscuration hardly explains the large line emission observed at minimum and the long brightness decay after maximum whereas variations of the red giant brightness imply unrealistic changes in the stellar radius. \\item At present, the available data for AE\\,Cir are hard to reconciliate with canonical models for symbiotic outbursts. %\\item {\\bf The similar visibility of the secondary star at the low and high state probably rules out the %hypothesis of the hot component outbursts and suggests an scenario where the %outbursts are due to thermal instabilities in an accretion disc fed by a Roche lobe filling giant star. %In this case the low state spectrum is dominated by the giant star and emission lines arising from wind accretion. %On the contrary, the hot state is characterized by a blue continuum arising from the hot and luminous disc in outburst %and a diminishing of the spectral features arising from the accreted wind. Irradiation of the secondary star by the %outbursting disc, and eventually stellar pulsation, could explain its variable brightness. The disappearance of the $\\lambda$ %6830 \\AA~ feature at the high state could be explained if the O\\,VI photon forming region is occulted by the accretion disc %during outburst.} \\end{itemize}" }, "0710/0710.2317_arXiv.txt": { "abstract": "The goal of this research is to investigate how well various turbulence models can describe physical properties of the upper convective boundary layer of the Sun. An accurate modeling of the turbulence motions is necessary for understanding the excitation mechanisms of solar oscillation modes. We have carried out realistic numerical simulations using several different physical Large Eddy Simulation (LES) models (Hyperviscosity approach, Smagorinsky, and dynamic models) to investigate how the differences in turbulence modeling affect the damping and excitation of the oscillations and their spectral properties and compare with observations. We have first calculated the oscillation power spectra of radial and non-radial modes supported by the computational box with the different turbulence models. Then we have calculated the work integral input to the modes to estimate the influence of the turbulence model on the depth and strength of the oscillation sources. We have compared these results with previous studies and with the observed properties of solar oscillations. We find that the dynamic turbulence model provides the best agreement with the helioseismic observations. ", "introduction": "Dominant acoustic sources within the Sun are generated by strong fluctuations in the outer convective layers. Turbulent motions stochastically excite the resonant modes via Reynolds stresses and entropy fluctuations. The dominant driving comes from the interaction of the nonadiabatic, incoherent pressure fluctuations with the coherent mode displacement \\citep{Nordlund2001}. The modes excitation sources occur close to the surface, mainly in the intergranular lanes and near the boundaries of granules \\citep{Stein2001}. Thus an accurate modeling of the turbulence motions is necessary to understand the excitation mechanisms of solar oscillation modes. The correct choice of turbulence model is also important in many other astrophysical simulations. The objective of this research is to study the influence of turbulence models on the excitation mechanisms by means of realistic numerical simulations. We have compared different physical Large Eddy Simulation (LES) models (Hyperviscosity approach, Smagorinsky, and dynamic models) to show the influence on the damping and excitation of the oscillations. The organization of this paper is as follows. In \\S2, we describe the main lines of the code and the different turbulence models. The kinetic energy of the radial modes obtained with the different turbulence models are presented in \\S 3. Then a comparison of the results obtained with the different turbulence models for the non-radial modes is given in \\S 4. The work integral input to the modes is calculated in \\S5 in order to estimate the influence of the turbulence models on the depth of the oscillation sources. ", "conclusions": "The goals of this research was to investigate how well various turbulence models can describe the convective properties of the upper boundary layer of the Sun and to study the excitation and damping of acoustic oscillations. Results obtained with the hyperviscosity approach have been compared with those obtained with the Smagorinsky and dynamic turbulence models. We have seen that the dissipation is very high with the Smagorinsky model while the hyperviscosity approach and dynamic modes give similar results. Besides we find that the dynamic turbulence model provides the best agreement with observations." }, "0710/0710.5701_arXiv.txt": { "abstract": "{} {We present the period analysis of unfiltered photometric observations of V5116~Sgr (Nova Sgr 2005 \\#2) and we search for superhump candidates in novae remnants.} {The PDM method for period analysis is used. The masses of the novae componets are estimated from the secondary mass -- orbital period and primary mass -- decline time relations.} {We found that 13 nights of V5116~Sgr observations in the year 2006 are modulated with a period of $0.1238 \\pm 0.0001$~d ($2.9712 \\pm 0.0024$~h). Following the shape of the phased light curves and no apparent change in the value of the periodicity in different subsamples of the data, we interpret the period as orbital in nature. The binary system then falls within the period gap of the orbital period distribution of cataclysmic variables. From the maximum magnitude -- rate of decline relation, we estimate the maximum absolute visual magnitude of $M_{\\rm Vmax} = -8.85 \\pm 0.04$~mag using the measured value of decline $t_{\\rm 2} = 6.5 \\pm 1.0$~d. The mass-period relation for cataclysmic variables yields a secondary mass estimate of about $0.26 \\pm 0.05~{\\rm M}_{\\rm \\odot}$. We propose that V5116~Sgr is a high inclination system showing an irradiation effect of the secondary star. No fully developed accretion disc up to the tidal radius with the value lower than $3.5~10^{10}$~cm is probable. The mass ratio was estimated in a few novae and the presence or absence of superhumps in these systems was compared with the mass ratio limit for superhumps of about 0.35. We found that in the majority of novae with expected superhumps, this variability has not been found yet. Therefore, more observations of these systems is encouraged.} {} ", "introduction": "Novae are a subclass of cataclysmic variable stars. In these interracting binaries, the white dwarf is accreting the matter transfered from the secondary star. The accretion disc may be formed in the non magnetic case. The intermediate polar systems have a truncated disc and polar systems have magnetic field strong enough to prevent the disc formation (see Warner 1995 for review). The accumulation of critical amount of accreted material onto the white dwarf surface results in a nova explosion. The distribution of orbital periods in cataclysmic variables shows a period gap between about 2 and 3 hours. Novae do not show this lack of objects in the mentioned interval. V5116~Sgr (Nova Sgr 2005 \\#2) was discovered by Liller (2005) on 2005 July 4.049 UT. The nova had a magnitude $\\sim 8.0$ on two red photographs. An unfiltered CCD image from 2005 July 5.085 UT showed the object at mag 7.2. The spectrum from 2005 July 5.099 UT showed H$_{\\rm \\alpha}$ with the FWHM of $\\sim 970$ km~s$^{-1}$. The expansion velocity derived from the sharp P~Cyg profile was $\\sim 1300$ km~s$^{-1}$. The position of the nova was measured by Gilmore and Kilmartin (2005) and Jacques (2005). Gilmore and Kilmartin (2005) searched for the nova precursor, but no convincing candidate has been found. Russell et al. (2005) performed a 0.8 -- 2.5 $\\mu$m spectroscopy of the nova on 2005 July 15. The object showed emission lines of H~I, He~I, C~I, N~I, Ca~II and O~I with a FWHM $\\sim$ 2200 km~s$^{-1}$. He~I showed P~Cyg profile at 1.0830 and 2.0581 $\\mu$m. No thermal dust emission was observed. After the nova explosion the accretions disc is destroyed. The new disc is reformed by the stream of matter flowing from the secondary, interracting with itself and forming a ring in the circularisation radius. The disc starts to form by the viscous shearing. Matter losing angular momentum is moving invard and the excess of angular momentum is transported by the matter flowing outward. The invard moving matter touchs the white dwarf and the disc is reformed (Pringle 1981). If the white dwarf is magnetic, the matter interracts with the magnetosphere and is then conducted by the magnetic field to the magnetic poles. The interraction of the flow with the poles of the rotating star is observed as periodic signal with the spin period of the white dwarf. In the case of intermediate polars the spin period is usually much shorter than the orbital period (Patterson 1994, Hellier 1996). The polar systems without a disc are synchronous rotators, hence the spin period is equal or close to the orbital period (see e.g. Schmidt and Stockman 1991). Nova V1500 Cyg (polar system without a disc) changed the period from 0.141~d to 0.138~d and then stabilised at 0.140~d (Patterson 1978, 1979). The difference of 1.8\\% between the rotation period of the white dwarf and the orbital period is ascribed to the effects of the nova explosion in 1975. The synchronisation can be then corrupted by the nova explosion, but the spin period remain still very close to the orbital period. This is in contrast with intermediate polars with disc presence. Searching for periods in novae allows to study the orbital distrubution and evolution of these systems. Currently, there are about 50 novae with known orbital periods (Warner 2002) with typical values ranging from 2 to 9 hours. The existence of the accretion disc or its renovation after the nova explosion is confirmed by the detection of the superhump period (see e.g. Retter et al. 1997, Kang et al. 2006a) or the spin period of a magnetic white dwarf in the case of intermediate polars (see e.g. Retter et al. 1998). Superhumps are caused by a precessing accretion disc generally in systems with mass ratio $M_{\\rm 2}/M_{\\rm 1} < 0.35 \\pm 0.02$ (see Patterson et al. 2005 for review), in which the disc radius reaches the 3:1 resonance. The mass ratio indicates that systems with massive primaries and low mass secondaries are probable superhumpers. Novae in general possess high mass white dwarfs (see e.g. Warner 1995, Smith and Dhillon 1998, Webbink 1990) and short orbital periods sugest low mass secondaries. It is therefore meaningfull to search for superhump variability in novae with short orbital period. We have an ongoing program to observe novae with small telescopes to search for periodicities in their optical light curves. In this paper we report the detection of one periodicity in our photometric data ($P = 0.1238 \\pm 0.0001$~d) of V5116~Sgr and we discuss the presence of superhumps in nova remnants. In Section~\\ref{obs} we present our observational material. The long-term light curve with the period analysis of the data is presented in Sec.~\\ref{data_anal} and in Sec.~\\ref{disc} we discuss the long-term light curve behaviour ({Sec.~\\ref{disc_v5116_l}}), the results of the period analysis (Sec.~\\ref{disc_v5116_s}) and superhump search in nova remnants (Sec.~\\ref{disc_SH}). ", "conclusions": "\\label{disc} \\subsection{Long-term variations of V5116~Sgr} \\label{disc_v5116_l} The long-term light curve of the nova V5116~Sgr (Fig.~\\ref{curve} -- bottom panel) shows a transition from smooth decline to probable oscillations. Several models has been suggested for such ``transient'' phase. One of them explains the transition as the time when the accretion disc is re-established (Retter 2002). The author propose a connection between this phase and intermediate polars. There is no indication of magnetic nature of the white dwarf in V5116~Sgr yet. Spin period of the white dwarf indicating the intermediate polar type of this ``transient'' phase in novae was detected for example in V4745~Sgr by optical observations (Dobrotka et al. 2006a), V1494~Aql (Drake et al. 2003) and V4743~Sgr by X-ray observations (Ness et al. 2003). If this ``transient'' phase interpretation of Retter (2002) is applicable in the case of V5116~Sgr, the accretion must be then re-established and the accretion disc should be present. \\subsection{Short-term variations of V5116~Sgr} \\label{disc_v5116_s} We have identified one periodicity in the light curve of the nova V5116~Sgr about 15 months after the maximum brightness. The value is $P = 0.1238 \\pm 0.0001$~d ($2.9712 \\pm 0.0024$~h). The upper limits of the period difference between two subset of data with $\\simeq 16$ days of mean distance analysed in Fig.~\\ref{power2} is $1.5~10^{-4}$~d which gives $|\\dot{P}| \\simeq 0.94~10^{-5}$. Following Patterson et al. (1993) the mean variation of the superhump period in SU~UMa superhumping sytems is $|\\dot{P}| \\simeq 3-10~10^{-5}$ (see their Table.~1). For the recurrent nova VY~Aqr the authors derived a variation of $|\\dot{P}| \\simeq 8.2~10^{-5}$. Our value of $|\\dot{P}|$ comes probably from the errors of period measurements rather than from real period changes. The period seems then to be constant during our observations and the shape of the folded light curve suggests a primary and a secondary eclipse. We therefore propose that this periodicity is the orbital period of the binary system. Such a period is at the lower end of the mostly populated region of orbital periods in novae (Warner 2002). The dominant first harmonic frequency in the power spectra is a result of the clear structure primary -- secondary eclipse in the folded light curve which suggests a high inclination angle of the binary system. Using the orbital period of the system and equation (9) from Smith and Dhillon (1998) we obtain a rough estimate for the secondary star mass of $0.26 \\pm 0.05$~M$_{\\rm \\odot}$. Using a mean white dwarf mass of $0.85 \\pm 0.05$~M$_{\\rm \\odot}$ from Smith and Dhillon (1998), we find a mass ratio (secondary/primary mass) of $M_{\\rm 2}/M_{\\rm 1} = 0.3 \\pm 0.1$. After the nova explosion, the hot white dwarf may heat and irradiate the cooler companion. The observed orbital light curve of the nova can be the result of the aspect variations or eclipses of the secondary due to heating from the hot primary and the asymmetry in the pulse profiles could be produced once the shape of the secondary is of a tear drop model. The irradiation effects in classical novae can also be detected long after the outburst stage (e.g., V1500~Cyg; Sommers and Naylor 1999, DN~Gem; Retter et al. 1999). Two eclipse like features are present in the folded light curve of V5116~Sgr (Fig.~\\ref{folded}). The shape is similar to the light curve of V2540~Oph in 2003 (Ak et al. 2005) with a dip at phase $\\sim 0.5$ ($\\sim 33\\%$ amplitude of the primary minimum for V2540~Oph and $\\sim 70\\%$ for V5116~Sgr). The authors concluded that V2540~Oph is likely a high inclination system either showing an irradiation effect or having a spiral structure in its accretion disc. Woudt and Warner (2003a) noted that one of the following requirements must be fulfilled for the large amplitude orbital modulations to be seen in the light curve of a recent nova in which the accretion disc does not dominate the luminosity of the system: 1) the disc is foreshortened but the secondary is seen (a high inclination angle), 2) the disc has small dimensions, 3) no disc (case of polars). In the case of V5116~Sgr the polar interpretation is possible because of the orbital period distribution of polars which peaks below 5 hours (Warner 1995), but nothing else indicate this option. The small dimension of the disc is supported by; 1) the short orbital period, 2) the post nova stage when the disc is reforming after explosion. A comparison to two novae within the period gap IM~Nor and DD~Cir with present irradiation effect (Woudt and Warner 2003b,c) can be made. The light curve of V5116~Sgr is different from those of IM~Nor and DD~Cir but similar to V2540~Oph. IM~Nor and DD~Cir showed very small dips at phase 0.5 interpretted as partial eclipses of the irradiated secondary by the disc or matter stream. The light curve shape depends on the disc radius and on the inclination angle. The deep secondary eclipses in V5116~Sgr require a large disc or a high inclination. The strength of the irradiation effect (white dwarf post-outburst temperature) can also play a significant role. The differences in the phase of the maxima in Fig.~\\ref{folded}b,c can be due to spiral structures in the disc as mentioned before in regard with the nova V2540~Oph. The reconstruction of an accretion disc after the nova eruption is indicated by the discovery of the white dwarf spin or by the superhump period. A probable spin period of the white dwarf was detected $\\sim 1$ year after the outburst in V4745~Sgr (Dobrotka et al. 2006a), $\\sim 2.75$ years in V4743~Sgr (Kang et al. 2006b) and $\\sim 15$ months after the maximum in V1425~Aql (Retter et al. 1998). Several systems show superhumps as early as two and a half months after the outburst like V4633~Sgr (a spin period is another option in this case, Lipkin et al. 2001) or two years after the outburst like V1974~Cyg (Retter et al. 1997). In our V5116~Sgr light curves extending $\\sim 470$ days after outburst we did not find any photometric variations consistent with white dwarf spin modulation or superhump properties. We can not say anything about the magnetic nature of the white dwarf, but the components mass ratio indicates that the superhump existence is probable. The mass ratio using the primary mass average is not decisive. Using the derived secondary mass 0.26~M$_{\\rm \\odot}$ (Section~\\ref{disc_v5116_s}) the mass ratio is lower than 0.35 for primary mass higher than $0.74$~M$_{\\rm \\odot}$. Taking all known primary masses in cataclysmic variables (Ritter and Kolb 2003), 63\\% have higher or equal mass than $0.74$~M$_{\\rm \\odot}$. This probability value is not enought to make sure that superhumps are expected inV5116~Sgr, but the search for such variability could be fruitful. However V5116~Sgr is a very fast nova. According to equation (13) from Livio (1992) and taking $t_{\\rm 3} = 20.2 \\pm 1.9$~d derived in this paper (Section~\\ref{t2t3}) we obtain a mass estimate of the white dwarf of $1.04 \\pm 0.02$~M$_{\\rm \\odot}$ and thus $M_{\\rm 2}/M_{\\rm 1} = 0.25 \\pm 0.05$. The presence of superhumps is then expected if a disc is fully developed up to tidal radius. The superhump period is a few percent longer or shorter than the orbital period. The possibility that the periodicity found in this paper is a superhump is rejected following the stability discussion and the shape of the folded light curve (Fig.~\\ref{folded}). The mean shape of superhumps is typically an asymetric sinusoid and our data show typical eclipse like features. The presence of the disc in cataclysmic variables depends on the mass loss from the secondary. The matterial supplied from the secondary within the period gap depends on the strength of the magnetic braking driving the secondary out of thermal equilibrium (mass loss time scale shorter than the thermal time scale). The stronger the braking, the wider the gap will be and the higher is the upper end of the period gap. If the magnetic braking is low enought, the mass loss time scale may never become shorter than the thermal time scale. In the case of novae another condition is important. The nova explosion heats the secondary which leads to an enhanced mass transfer. Therefore the complete absence of accretion discs in novae within the period gap is not expected. \\subsection{Searching for superhumps in other nova remnants} \\label{disc_SH} The orbital period 2.462~h of the mentionned IM~Nor (Woudt and Warner 2003b) yields a secondary mass estimate of 0.20~M$_{\\rm \\odot}$ following Smith and Dhillon (1998). The decay time $t_{\\rm 3} \\simeq 50$~d (Kato et al. 2002) yields a primary mass estimate of $\\simeq 0.86$~M$_{\\rm \\odot}$ (Livio 1992) giving a mass ratio $\\simeq 0.23$. Woudt and Warner (2003b) concluded that they did not observe the superhump modulation. This conclusion with the mass ratio safely lower than 0.35 indicate that the existence of a fully developed disc is not probable. DD~Cir (Woudt and Warner 2003c) has a period of 2.339~h yielding a secondary mass estimate of 0.18~M$_{\\rm \\odot}$. The decay time $t_{\\rm 2} \\simeq 4.5$~d (Liller 1999, no other estimates are available) place the object in the class of very fast novae (similar to V5116~Sgr). The decay time $t_{\\rm 3} \\simeq 10$~d following the equation $t_{\\rm 3} \\simeq 2.75~t_{\\rm 2}^{0.88}$ from Warner (1995) gives a rought white dwarf mass estimate of $\\simeq 1.16$~M$_{\\rm \\odot}$ (mass ratio $< 0.2$). Woudt and Warner(2003b) interpretted the deep eclipse as obscuration of the disc and did not detect any periodicity near the orbital period in the period analysis. Therefore, following the tidal instability model it is again probable that the disc is not fully developed up to 3:1 resonance radius 3 years after the outburst. However the authors argued that the disc radius is 47 \\% of the orbital separation. Using the third Kepler law we obtain a disc radius of $\\sim 3.1~10^{10}$~cm. The disc in DD~Cir is then large enough to reach the 3:1 resonance radius calculated from the equation (3.39) from Warner (1995); $r_{\\rm 3:1}\\sim 3~10^{10}$~cm. The result is marginally in contrast with the absence of superhumps. Taking the derived values for V5116~Sgr we get $r_{\\rm 3:1}\\simeq 3.5~10^{10}$~cm as an upper limit of the disc radius taking the absence of superhumps into consideration. The boundaries of the eclipse in our light curve are not so clear as in the DD~Cir case, therefore not suitable to estimate the disc radius. Comparison to other systems follows the same way as in the case of IM~Nor and DD~Cir. We took components masses from the Catalogue of Ritter and Kolb (2003) and other masses are estimated from the orbital period (using Smith and Dhillon 1998) and $t_{\\rm 3}$ time (using Livio 1992). We rejected systems with unknown or non measurable $t_{\\rm 2}$ or $t_{\\rm 3}$ time (peculiar light curve, strong oscilations in the decline) without catalogue mass values (no information on mass ratio, ex: RS~Car), magnetic novae (AM~Her systems without disc, ex: V1500~Cyg), systems with orbital period $> 10$ h (Smith and Dhillon fitting not adequate, systems with evolved secondary are then excluded too, ex: DI~Lac, V841~Oph, V723~Cas, GK~Per), systems with insuficient photometric observations to detailed period analysis (ex: V500~Aql, V446~Her, HZ~Pup, DY~Pup, CT~Ser). The final list of systems is in Table~\\ref{systems}. \\begin{figure} \\includegraphics[width=90mm]{q_p.eps} \\caption{Selected nova remnants in the mass ratio ($q$) -- orbital period ($P_{\\rm orb}$) plane. Open circle -- non detected superhumps, filled circle -- detected superhumps. Panel a -- mass ratio from combined sourcees (catalogue and estimates from $P_{\\rm orb}$, $t_{\\rm 3}$ time), panel b -- mass ratio only from catalogue values (Ritter and Kolb 2003). The dashed line is the 0.35 limit.} \\label{q_p} \\end{figure} The results are depicted in Fig.~\\ref{q_p}. The lower panel only shows the systems with known catalogue mass values and the upper panel is after the addition of estimated mass values in this work. Filled circles are systems with detected superhumps. The critical mass ratio of 0.35 is also shown. The results presented in the lower panel are as expected from the mass ratio limit for superhumps, but the statistical set is small. Adding other mass estimates changes the situation. 7 systems below or close to the limit 0.35 show superhumps but 14 systems do not. The estimated mass ratio has a mean difference of 0.2 from the catalogue values. The differences are distributed randomly. Therefore there is no reason to suspect that all mass ratios are systematically underestimated. The 7 systems with detected superhumps are safely located in comparing to the 0.35 critical value as expected by theory and are only a third of all systems falling below or close to the mass ratio of 0.35. The majority of all systems occupy the orbital period interval of 2 -- 4 hours. Following Patterson (2005) this interval has a $\\sim$ 40 -- 90 \\% probability to observe superhumps. 7 systems versus 14 from our investigation yields only $\\sim 30$ \\%. Superhumps are observed during the outbursts of short orbital period dwarf novae (SU~UMa stars) where the disc reaches the 3:1 resonance radius (superoutburst case). The disc during this active stage (outburst -- active and hot stage, quiescence -- not active and cold stage) reaches larger radius that in quiescence (see Lasota 2001 for review). The irradiation of the secondary by the central white dwarf in nova remnants can produce enhanced mass transfer rate strong enough to keep the disc in this stable hot active stage (as in the nova like stars) with the larger diameter (Hameury and Lasota 2005). In addition, superoutbursts in SU~UMa stars, when superhumps are observed, can be explained by the irradiation and the enhanced mass transfer (Smak 1995, 2004, Hameury et al. 2000, Schreiber et al. 2004). Therefore, permanent superhumpers are possible after a nova outburst. A few possible interpretation of the superhump lack can be: 1) not systematically applicable tidal instability theory; tidal torques are probably not the main and only condition of superhump existence as concluded by Hameury and Lasota (2005), 2) disc radius not developed up to tidal radius, 3) superhumps present but not detected, because of insufficient sets of data." }, "0710/0710.5192_arXiv.txt": { "abstract": "The group of 7 thermally emitting and radio-quiet isolated neutron stars (INSs) discovered by ROSAT constitutes a nearby population which locally appears to be as numerous as that of the classical radio pulsars. So far, attempts to enlarge this particular group of INSs finding more remote objects failed to confirm any candidate. We found in the 2XMMp catalogue a handful of sources with no catalogued counterparts and with X-ray spectra similar to those of the ROSAT discovered INSs, but seen at larger distances and thus undergoing higher interstellar absorptions. In order to rule out alternative identifications such as an AGN or a CV, we obtained deep ESO-VLT and SOAR optical imaging for the X-ray brightest candidates. We report here on the current status of our search and discuss the possible nature of our candidates. We focus particularly on the X-ray brightest source of our sample, 2XMM J104608.7-594306, observed serendipitously over more than four years by the XMM-Newton Observatory. A lower limit on the X-ray to optical flux ratio of $\\sim$ 300 together with a stable flux and soft X-ray spectrum make it the most promising thermally emitting INS candidate. Beyond the finding of new members, our study aims at constraining the space density of this population at large distances and at determining whether their apparently high local density is an anomaly or not. ", "introduction": "One of the outstanding results of ROSAT is the discovery of seven X-ray bright isolated neutron stars (INSs). These slowly rotating and radio-quiet neutron stars display thermal emission with $kT \\sim$ 40 -- 110 eV, undergo little interstellar absorptions and are not associated with any supernova remnant (see reviews in \\cite{tre00} and \\cite{hab07}). Several have identified faint optical counterparts with $B \\sim$ 25.8 -- 28.6. Proper motion studies (see Motch et al., these proceedings, and references therein) have shown that they are most probably young cooling neutron stars, with ages of a few 10$^5$ years. Their proximity and the apparent absence of strong non-thermal activity turn them into unique laboratories for testing radiative properties of neutron star surfaces in extreme conditions of gravitational and magnetic fields. Moreover, the possibility of measuring their distances through parallaxes \\cite{ker07} or from the distribution of absorption on the line of sight \\cite{pos07} can eventually bring important constraints on the debated equation of state of matter in neutron star interiors. In the solar neighborhood, ROSAT INSs are as numerous as young radio and $\\gamma$-ray pulsars. It is not clear whether this group is homogeneous. In particular, the absence of radio emission can be either due to the presence of intense magnetic fields, indeed inferred from the measured spin down rates and cyclotron X-ray spectral features, or due to the fact that the radio pencil beam, which narrows at long spin periods, does not sweep over the earth. Considering that ROSAT had not enough sensitivity and spatial resolution to detect the thermal emission of distant sources, the population of cooling radio-quiet INSs could represent a considerable fraction of the total neutron star population of the Galaxy, undetectable in radio surveys \\cite{mot07}. In any case, our knowledge of the overall population characteristics will remain highly unsatisfactory as long as only seven objects are known. \\begin{table} \\begin{tabular}{l c r r c r r} \\hline \\tablehead{1}{l}{t}{Source} & \\tablehead{1}{c}{t}{Cand} & \\tablehead{1}{r}{t}{RA} & \\tablehead{1}{r}{t}{DEC} & \\tablehead{1}{c}{t}{$R_{90}$} & \\tablehead{1}{r}{t}{Count Rate} & \\tablehead{1}{r}{t}{Mag}\\\\ & & (J2000) & (J2000) & arcsec & s$^{-1}$ & $R$\\\\ \\hline 2XMM J104608.7-594306 & 065 & 10 46 08.7 & -59 43 06.1 & 1.33 & 0.060(4) & $>$ 25\\\\ 2XMM J121017.0-464609 & 164 & 12 10 17.1 & -46 46 11.2 & 2.60 & 0.027(6) & 20.3 \\\\ 2XMM J010642.3+005032 & 318 & 01 06 42.4 & +00 50 31.3 & 3.80 & 0.020(5) & 24.5 \\\\ 2XMM J214026.1-233222 & 364 & 21 40 26.2 & -23 32 22.3 & 1.90 & 0.0181(20) & 23.8 \\\\ 2XMM J125904.5-040503 & 604 & 12 59 04.6 & -04 05 02.3 & 1.90 & 0.0129(21) & 21.2 \\\\ 2XMM J125045.7-233349 & 681 & 12 50 45.7 & -23 33 47.7 & 3.30 & 0.0122(21) & 22.1 \\\\ \\hline \\end{tabular} \\caption{INS candidates selected for optical investigation which have already been observed. Count rates are for the EPIC pn camera in the full XMM-Newton energy band (0.2--12.0 keV). We list the $R$ magnitude of the brightest object present in the error circles of the X-ray sources.\\label{tab_cand}} \\end{table} \\begin{figure} \\includegraphics[height=.28\\textheight]{besteso_0010.eps} \\caption{The positions of our INS candidates, selected from $\\sim$ 7.5$\\cdot$10$^4$ XMM sources, are shown in the HR$_1 \\times$ HR$_2$ diagram (squares). The known ROSAT INSs (crosses) occupy the lowest (less absorbed) part of the diagram. Dashed lines denote soft absorbed blackbodies of different temperatures (50-200 eV) and column absorptions. The nine INS candidates selected for optical follow-up during the present year are highlighted with circles. The X-ray brightest and most promising candidate (source 65) can be noticed by its somewhat smaller error bars. \\label{fig_bestcand}} \\end{figure} ", "conclusions": "Our search for new thermally-emitting INSs in the 2XMMp catalogue has revealed a handful of interesting and previously unknown soft X-ray sources among which source 65 is, by far, the most promising candidate. The analysis of its X-ray emission, although based on archival data obtained with non-optimal configurations, reveals an intrinsically soft energy distribution, apparently stable on long time scales. The derived $N_{\\textrm{H}}$ is consistent with that observed towards Eta Carinae and its cluster ($N_{\\textrm{H}} \\sim$ 3$\\cdot$10$^{21}$ cm$^{-2}$). Scaling from RX J0720.4$-$3125 yields a distance of $\\sim$ 3.9 kpc probably compatible with the 2.5 kpc assumed for the Carina Nebula \\cite{2007ApJ...656..462S}. Optical follow-up observations failed to reveal counterparts brighter than $R \\sim$ 25, supporting the idea that source 65 is a new thermally emitting INS. We have already obtained optical data for some of the other candidates. Preliminary analysis of these data reveals the presence of faint optical candidates within the error circles in all cases. We plan to use optical colour indices and spectra to identify these sources, thus characterizing this sample of soft objects strictly selected from more than 120 thousand entries present in the 2XMMp catalogue. \\begin{theacknowledgments} This work has been supported by Funda\\c c\\~ao de Amparo \\`a Pesquisa do Estado de S\\~ao Paulo (FAPESP), Coordena\\c c\\~ao de Aperfei\\c coamento de Pessoal de N\\'ivel Superior (CAPES), Brazil and by Universit\\'e Louis Pasteur, Strasbourg, France. \\end{theacknowledgments}" }, "0710/0710.3816.txt": { "abstract": "Spectral studies of quiescent emission and bursts of magnetar candidates using XMM-Newton, {\\it Chandra} and {\\it Swift} data are presented. Spectra of both the quiescent emission and the bursts for most magnetar candidates are reproduced by a photoelectrically absorbed two blackbody function (2BB). There is a strong correlation between lower and higher temperatures of 2BB ($kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$) for the magnetar candidates of which the spectra are well reproduced by 2BB. In addition, a square of radius for $kT_{\\mathrm{LT}}$ ($R_{\\mathrm{LT}}^2$) is well correlated with a square of radius for $kT_{\\mathrm{HT}}$ ($R_{\\mathrm{HT}}^2$). % A ratio $kT_{\\mathrm LT}/kT_{\\mathrm HT} \\approx 0.4$ is nearly constant irrespective of objects and/or emission types (i.e., the quiescent emission and the bursts). This would imply a common emission mechanism among the magnetar candidates. The relation between the quiescent emission and the bursts might be analogous to a relation between microflares and solar flares of the sun. Three AXPs (4U\\,0142$+$614, 1RXS\\,J170849.0$-$400910 and 1E\\,2259$+$586) seem to have an excess above $\\sim$\\,7\\,keV which well agrees with a non-thermal hard component discovered by INTEGRAL. ", "introduction": "Among peculiar celestial objects in the universe, a dense highly magnetized neutron star ($\\rho \\sim 10^{14}$\\,g\\,cm$^{-3}$ and $B \\sim 10^{15}$\\,G), so-called ``magnetar'' \\citep{duncan1992, paczynski1992, thompson1995, thompson1996}, would be one of the most exotic objects. % Soft gamma repeaters (SGRs) and anomalous X-ray pulsars (AXPs) are well known as magnetar candidates. An apparent difference between the SGRs and the AXPs would be considered from their first detections. The SGRs were discovered as sporadically bursting objects, while the AXPs were regarded as peculiar pulsars with long spin periods. However, current observations unveil a lot of similarities between these objects. They have, for instance, long spin periods ($P \\sim $ 5-12\\,s) with spindown rates of $\\dot{P} \\sim 10^{-10}$-$10^{-13}$\\,s\\,s$^{-1}$, no signature of a companion star, a distribution around the galactic plane (two magnetar candidates are in other galaxies), quiescent soft X-ray emission. Several of these objects have non-thermal hard ($>20$\\,keV) components, some are associated with supernova remnants (SNRs), and bursting activity is not confined to the SGRs but is observed in the AXPs as well. Considering these similarities, the SGRs and the AXPs should be classified into a common class of objects. So far, five SGRs (0501$+$4516, 0526$-$66, 1627$-$41, 1806$-$20 and 1900$+$14) are known \\citep{woods2006, barthelmy2008} as well as three candidates, SGR\\,1801$-$23 \\citep{cline2000}, SGR\\,1808$-$20 \\citep{lamb2003} and SGR/GRB\\,050925. SGR/GRB\\,050925 was regarded as a gamma-ray burst (GRB) when first detected, but soon after was recognized as a new SGR \\citep{holland2005}. On the other hand, ten AXPs (1E\\,2259$+$586, 1E\\,1048.1$-$5937, 4U\\,0142$+$614, 1RXS\\,J170849.0$-$400910, 1E\\,1841$-$045, XTE\\,J1810$-$197, AX\\,J1845$-$0258, CXOU\\,J010043.1$-$721134, CXOU\\,J164710.2$-$455216 and 1E\\,1547.0$-$5408) are known to date \\citep{woods2006, dib2008} with one AXP candidate, AXP\\,CXOU\\,J160103.1$-$513353 \\citep{park2006}. % A short burst from AXP\\,CXOU\\,J164710.2$-$455216 was detected by {\\it Swift} BAT \\citep{krimm2006} at 01:34:52 on 2006 September 21. The follow-up observations performed by {\\it Swift} XRT found a remarkable result in which the quiescent emission of post-burst became 190 times brighter than that of pre-burst \\citep{campana2006}. In addition to these objects, AX\\,J1818.8$-$1559 discovered by ASCA \\citep{sugizaki2001} recently exhibited a short burst \\citep{mereghetti2007b} similar to those from the magnetar candidates. Therefore AX\\,J1818.8$-$1559 could be a new SGR or AXP \\citep{mereghetti2007b}. The most exciting phenomena among the magnetar candidates would be a sudden release of huge energy in rather short period, the so-called {\\it giant flares} from the SGRs. They typically have a short intense spike which last less than 1\\,s, and followed by a long pulsating tail which lasts a few hundred seconds. Their peak energy flux can be larger than $\\sim10^{6}$ times Eddington luminosity. Theoretical studies suggested that the giant flares were triggered by a catastrophic deformation of the neutron star crust due to a torsion of the strong magnetic field (e.g., \\cite{thompson2001}). Some different emission mechanisms have been proposed by several authors \\citep{yamazaki2005, lyutikov2006, cea2006}. % In the past three decades, three giant flares were recorded. The first detection, from the source now known as SGR\\,0526$-$66 in Large Magellanic Cloud (LMC), was made on March 5 in 1979 \\citep{mazets1979, cline1980, evans1980, fenimore1996}. % The second one from SGR\\,1900$+$14 was recorded on August 27 in 1998 \\citep{hurley1999b, feroci1999, mazets1999, feroci2001, tanaka2007}. More recently, the most energetic giant flare from SGR\\,1806$-$20 was observed on December 27 in 2004 \\citep{cameron2005, gaensler2005, hurley2005, mazets2005, palmer2005, terasawa2005, tanaka2007}. The fluence of its initial intense spike with 600\\,ms was evaluated to be $\\sim 2$\\,erg\\,cm$^{-2}$ by the plasma particle detectors on the Geotail space probe \\citep{terasawa2005}. Soft X-ray spectra of the quiescent emission of the SGRs and the AXPs were observed by a number of satellites. Although their spectral model is still under discussion, two two-component models are proposed. % One of them is a photoelectrically absorbed two blackbody function (2BB). Spectral parameters of 2BB are reported by some authors for the SGRs \\citep{mereghetti2006a} and the AXPs \\citep{tiengo2002, morii2003, gotthelf2004, gotthelf2005, halpern2005, tiengo2005, israel2006, gotthelf2007}. % Typical lower and higher temperatures are $\\sim 0.5$\\,keV and $\\sim 1.4$\\,keV, respectively. % The other model is a photoelectrically absorbed power law plus a blackbody (PL$+$BB). Some authors report spectral parameters of PL$+$BB for the SGRs \\citep{marsden2001, kurkarni2003, mereghetti2005, mereghetti2006a, mereghetti2007a} and the AXPs \\citep{morii2003, patel2003, rea2003, gotthelf2004, mereghetti2004, woods2004, tiengo2005, gavriil2006, israel2006}. % A typical power law index and a blackbody temperature are $\\sim 3$ and $\\sim 0.5$\\,keV. % At present it is still unclear which model is more reliable or physically suitable. Recent observations by the International Gamma-Ray Astrophysics Laboratory (INTEGRAL) discovered a non-thermal hard component in the spectra of the quiescent emission above 20\\,keV for 5 magnetar candidates \\citep{molkov2005, gotz2006b, kuiper2006}. The non-thermal hard component is well reproduced by a power law model, $E^{-\\Gamma}$, where $\\Gamma$ ranges from 1.0-1.8, while the soft X-ray emission below $\\sim12$\\,keV, mentioned above, clearly indicates steeper power-law index of $\\sim 3$ if the PL$+$BB is applied as the model spectrum. Hence, the non-thermal hard emission seen by INTEGRAL is a different component and presumably has a different origin than the soft X-ray emission. Since some magnetar candidates have two different emission mechanisms, there seems to be more complex physics than expected before. % Moreover, the non-thermal hard component shows pulsations for three AXPs, 1RXS\\,J170849.0$-$400910, 4U\\,0142$+$614 and 1E\\,1841$-$045, through the INTEGRAL and RXTE observations \\citep{kuiper2006}, which is related to a neutron star rotation, and hence there are particle acceleration processes in the vicinity of neutron stars \\citep{kuiper2006}. If the energy source of the quiescent emission and the % short bursts is the magnetic field as thought to be, at least very similar physical process would govern both of them and their spectra could emerge alike. It is claimed based on High Energy Transient Explorer 2 (HETE-2) data that the most acceptable spectral model of the short bursts from two SGRs 1806\u00a1\u00dd20 and 1900+14 is 2BB even though it should be regarded just as an empirical model (Nakagawa et al. 2007). It would be also preferred to represent spectra of quiescent emissions by 2BB rather than BB+PL for SGRs, and even for AXPs if it is the same class of object. In this paper, we present a comprehensive spectral study with 2BB for both the quiescent emission and the for the magnetar candidates. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "The spectral studies using the photoelectrically absorbed two blackbody function (2BB) were presented for the quiescent emission and the burst of the magnetar candidates. % The spectra of the quiescent emission were well reproduced by a 2BB with some exceptions. The spectra of three AXPs (4U\\,0142$+$614, 1RXS\\,J170849.0$-$400910 and 1E\\,2259$+$586) seem to have an excess which might be due to a non-thermal hard component discovered by INTEGRAL. % The spectrum of the burst from the SGR candidate SGR\\,2013$+$34 was also well reproduced by 2BB. A strong linear correlations between $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ was found using 2BB spectra. The ratio $kT_{\\mathrm{LT}}/kT_{\\mathrm{HT}} \\sim 0.4$ is almost constant irrespectively of the objects and/or emission types (burst or quiescent emission). The relationship between $R_{\\mathrm{LT}}^2$ and $R_{\\mathrm{HT}}^2$ seems to have a linear correlation. Considering these correlations, there seems to be a common emission mechanism among these objects, and between the quiescent emission and the burst. The relationship between the quiescent emission and the burst might be similar to the relationship between microflares and an ordinary solar flares of the sun. The quiescent emission might be due to very frequent small activity similar to the microflares. On the other hand, the burst might be due to a relatively large activity similar to the ordinary solar flare. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Acknowledgement \\bigskip We would like to thank an anonymous referee for helpful comments and suggestions to improve our paper. This work is based on observations obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. % We would like to thank public data archive of {\\it Chandra}. % This research has made use of software provided by the Chandra X-ray Center (CXC) in the application packages CIAO, ChIPS, and Sherpa. % We acknowledge the use of public data from the Swift data archive. % YEN is supported by the JSPS Research Fellowships for Young Scientists. This work is supported in part by a special postdoctoral researchers program in RIKEN. %\\appendix %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Figures %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\\clearpage % \\begin{figure} \\begin{center} \\FigureFile(80mm,129mm){figure1.eps} \\end{center} \\caption{Time history of the (a) flux in 2-10 keV in units of ergs cm$^{-2}$ s$^{-1}$, (b) photoelectric absorption $N_{\\mathrm{H}}$ in units of cm$^{-2}$, (c) temperature of the lower blackbody $kT_{\\mathrm{LT}}$ in units of keV, (d) radius of the lower blackbody $R_{\\mathrm{LT}}$ in units of km, (e) temperature of the higher blackbody $kT_{\\mathrm{HT}}$ in units of keV and (f) radius of the higher blackbody $R_{\\mathrm{HT}}$ in units of km for the emissions of AXP\\,CXOU\\,J164710.2$-$455216 observed by XRT/{\\it Swift}.}\\label{fig:lc_persistent_cxou_j16} \\end{figure} \\begin{figure} \\begin{center} \\FigureFile(80mm,80mm){figure2.eps} \\end{center} \\caption{A schematic view of $\\nu{F}_{\\nu}$ spectra using 2BB$+$PL (a) and BB$+$2PL (b) for AXP\\,4U\\,0142$+$614. Spectral parameters of X-ray spectra (i.e., $\\lesssim$12\\,keV) are derived from our analyses using the XMM-Newton observation of 0112781101, while those of the non-thermal hard component (i.e., $\\gtrsim$20\\,keV) is obtained by INTEGRAL observations \\citep{kuiper2006}. The circles denote data derived from our analyses using the XMM-Newton observation of 0112781101. The squares represent INTEGRAL observations taken from Fig.7 in \\citet{kuiper2006} by eye. The dashed, dot-dash, dotted lines in (a) show the $kT_{\\mathrm{LT}}$ component, the $kT_{\\mathrm{HT}}$ component and the PL component for the hard spectrum, respectively. Those lines in (b) show the $kT$ component, the PL component for an X-ray spectrum and the PL component for the hard spectrum, respectively.}\\label{fig:eeufspec_summary} \\end{figure} \\begin{figure} \\begin{center} \\FigureFile(80mm,83mm){figure3.eps} \\end{center} \\caption{Relationship between the 2BB temperatures $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$. The triangles and squares denote the previous work on the bursts \\citep{feroci2004, olive2004, gotz2006a, nakagawa2007} and the quiescent emission \\citep{morii2003, gotthelf2004, gotthelf2005, tiengo2005, mereghetti2006a}, respectively. The circles and stars denote our work on the bursts and the quiescent emission, respectively. The line represents the best-fit power law model.}\\label{fig:bb_relation} \\end{figure} \\begin{figure} \\begin{center} \\FigureFile(80mm,80mm){figure4.eps} \\end{center} \\caption{Relationship between the square of the blackbody radii $R^2_{\\mathrm{LT}}$ and $R^2_{\\mathrm{HT}}$. The triangles and squares denote the previous work on the the bursts \\citep{olive2004, nakagawa2007} and the quiescent emission \\citep{morii2003, tiengo2005, mereghetti2006a}, respectively. The stars denote our work on the quiescent emission. The solid line shows a ratio of $R_{\\mathrm{HT}}^{2}$ to $R_{\\mathrm{LT}}^{2}$ of 0.01.}\\label{fig:radius_relation} \\end{figure} \\clearpage \\begin{figure} \\begin{center} \\FigureFile(160mm,80mm){figure5.eps} \\end{center} \\caption{{\\it Left:} Relationship between the lower temperature of 2BB $kT_{\\mathrm{LT}}$ and the square of the blackbody radii of 2BB $R_{\\mathrm{LT}}^2$. {\\it Right:} Relationship between the higher temperature of 2BB $kT_{\\mathrm{HT}}$ and the square of the blackbody radii $R_{\\mathrm{HT}}^2$. The dotted and dashed lines correspond to bolometric fluences of $10^{-8}$ and $10^{-9}$\\,ergs\\,cm$^{-2}$, respectively.}\\label{fig:kt_rsquared_relation} \\end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Tables %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\clearpage \\begin{table*} \\small \\caption{A summary of magnetar candidates which were employed in our study.}\\label{tab:magnetar_list} \\begin{center} \\begin{tabular}{lllll} \\hline\\hline Object\\footnotemark[$*$] & Satellite/Instrument\\footnotemark[$\\dagger$] & Period\\footnotemark[$\\ddagger$] & Distance\\footnotemark[$\\S$] & Ref.\\footnotemark[$\\|$] \\\\ \\hline SGR\\,0526$-$66 & {\\it Chandra} ACIS & 2000-2001 & 50 & (1) \\\\ SGR\\,1627$-$41 & XMM-Newton EPIC, {\\it Chandra} ACIS & 2001-2005 & 11 & (2), (3), (4) \\\\ SGR\\,1806$-$20 & XMM-Newton EPIC, {\\it Chandra} ACIS & 2000-2005 & 15\\footnotemark[$\\#$] & (5), (6) \\\\ SGR\\,2013$+$34 & {\\it Swift} BAT & 2005 & 10 & (7), (8) \\\\ SGR\\,1819$-$16 & XMM-Newton EPIC & 2003 & 10 & \\\\ AXP\\,CXOU\\,J010043.1$-$721134 & XMM-Newton EPIC, {\\it Chandra} ACIS & 2000-2005 & 57 & (9), (10) \\\\ AXP\\,4U\\,0142$+$614 & XMM-Newton EPIC & 2002-2004 & 3 & (11) \\\\ AXP\\,CXOU\\,J164710.2$-$455216 & {\\it Swift} BAT, {\\it Swift} XRT, {\\it Chandra} ACIS & 2005-2007 & 5 & (12), (13), (14), (15) \\\\ AXP\\,1RXS\\,J170849.0$-$400910 & XMM-Newton EPIC & 2003 & 5 & (16), (17) \\\\ AXP\\,1E\\,1841$-$045 & XMM-Newton EPIC & 2002 & 7\\footnotemark[$**$] & \\\\ AXP\\,1E\\,2259$+$586 & XMM-Newton EPIC & 2002-2005 & 3 & (18) \\\\ \\hline \\multicolumn{5}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of magnetar candidates (SGR\\,2013$+$34 denotes SGR candidate SGR/GRB\\,050925). \\par\\noindent \\footnotemark[$\\dagger$] Instrument and satellite names from which obtained the data used in our analysis. \\par\\noindent \\footnotemark[$\\ddagger$] Interval which these observations were performed. \\par\\noindent \\footnotemark[$\\S$] Distances to each object in units of kpc (see \\cite{woods2006} and references there in). \\par\\noindent \\footnotemark[$\\|$] (1) \\citet{kurkarni2003}; (2) \\citet{kouveliotou2003}; (3) \\citet{wachter2004}; (4) \\citet{mereghetti2006b}; (5) \\citet{kaplan2002}; (6) \\citet{mereghetti2005}; (7) \\citet{holland2005}; (8) \\citet{markwardt2005}; (9) \\citet{lamb2002}; (10) \\citet{majid2004}; (11) \\citet{guver2007}; (12) \\citet{krimm2006}; (13) \\citet{campana2006}; (14) \\citet{muno2006}; (15) \\citet{muno2007}; (16) \\citet{oosterbroek2004}; (17) \\citet{rea2007}; (18) \\citet{woods2004} \\par\\noindent \\footnotemark[$\\#$] The latest distance estimate is $d = 8.7$\\,kpc \\citep{bibby2008}. \\par\\noindent \\footnotemark[$**$] The latest distance estimate is $d = 8.5$\\,kpc \\citep{tian2008}. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\scriptsize \\caption{XMM-Newton observations of the quiescent emissions of the SGRs.}\\label{tab:sgr_obs_xmm} \\begin{center} \\begin{tabular}{lllllllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & \\multicolumn{2}{l}{Observation Date (MJD)\\footnotemark[$\\ddagger$]} & \\multicolumn{3}{l}{Observation Mode\\footnotemark[$\\S$]} & \\multicolumn{3}{l}{Exposure Time (ks)\\footnotemark[$\\|$]} & \\multicolumn{3}{l}{Source/Background Radii} \\\\ & & Start & End & pn & MOS1 & MOS2 & pn & MOS1 & MOS2 & pn & MOS1 & MOS2 \\\\ \\hline 1627$-$41 & 0204500201 & 53051.590 & 53051.992 & Full & Full & Full & 15.89 & 20.03 & 20.17 & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ \\\\ 1627$-$41 & 0204500301 & 53252.750 & 53253.131 & Full & Full & Full & 27.06 & 32.00 & 32.10 & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ \\\\ 1627$-$41 & 0202560101 & 53270.677 & 53271.281 & Small & P-W2 & P-W2 & 26.66 & 36.19 & 49.38 & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ & $\\timeform{10''}$/$\\timeform{10''}$ \\\\ 1806$-$20 & 0148210101 & 52732.566 & 52733.209 & Full & P-W3 & P-W3 & 4.84 & 5.65 & 5.62 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 1806$-$20 & 0148210401 & 52919.404 & 52919.663 & Full & P-W3 & P-W3 & 7.61 & 7.07 & 7.23 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 1806$-$20 & 0205350101 & 53254.377 & 53254.978 & Small & P-W3 & P-W3 & 30.21 & 39.14 & 39.56 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 1806$-$20 & 0164561101 & 53284.706 & 53284.925 & Small & Fast-U & Fast-U & 11.54 & $t$ & $t$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 1806$-$20 & 0164561301 & 53436.348 & 53436.636 & Small & Fast-U & Full & 7.37 & $t$ & 5.68 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 1806$-$20 & 0164561401 & 53647.427 & 53647.809 & Small & Fast-U & Full & 22.11 & $t$ & 28.68 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 2013$+$34 & 0212481201 & 53655.026 & 53655.334 & Full & Full & Full & 22.18 & 25.27 & 25.27 & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 1819$-$16 & 0152834501 & 52726.191 & 52726.310 & Full & Full & Full & 3.3 & 4.3 & 4.8 & $\\timeform{32''}$/$\\timeform{80''}$\\footnotemark[$\\#$] & $\\timeform{28''}$/$\\timeform{70''}$\\footnotemark[$\\#$] & $\\timeform{28''}$/$\\timeform{70''}$\\footnotemark[$\\#$] \\\\ \\hline \\multicolumn{13}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the SGRs (2013$+$34 denotes SGR candidate SGR/GRB\\,050925 and 1819$-$16 denotes SGR candidate AX\\,J1818.8$-$1559). \\par\\noindent \\footnotemark[$\\dagger$] XMM-Newton observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] Start and end time of observations. \\par\\noindent \\footnotemark[$\\S$] Observation mode for each instrument; full-window mode (Full), small-window mode (Small), partial-w2 mode (P-W2), partial-w3 mode (P-W3) and fast-uncompressed mode (Fast-U). \\par\\noindent \\footnotemark[$\\|$] Net exposure time for each instrument. $t$ denotes the data sets obtained by the MOS cameras in timing mode and not utilized. \\par\\noindent \\footnotemark[$\\#$] The background regions were extracted from an annular region whose center was the source position. The first values are source radii, and the inner radii of the background regions. The second values are outer radii of the background regions. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\small \\caption{{\\it Chandra} observations of the quiescent emissions of the SGRs.}\\label{tab:sgr_obs_cxo} \\begin{center} \\begin{tabular}{lllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & \\multicolumn{2}{l}{Observation Date (MJD)\\footnotemark[$\\ddagger$]} & Observation Mode\\footnotemark[$\\S$] & Exposure Time (ks)\\footnotemark[$\\|$] & Source/Background Radii \\\\ & & Start & End & & & \\\\ \\hline 0526$-$66 & 747 & 51547.017 & 51547.539 & FAINT & 39.86 & $\\timeform{1''}$/$\\timeform{1''}$ \\\\ 0526$-$66 & 1957 & 52152.937 & 52153.566 & FAINT & 48.45 & $\\timeform{1''}$/$\\timeform{1''}$ \\\\ 1627$-$41 & 1981 & 52182.205 & 52182.803 & FAINT & 48.93 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 1627$-$41 & 3877 & 52722.169 & 52722.494 & VFAINT & 25.67 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ \\hline \\multicolumn{7}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] SGR names. \\par\\noindent \\footnotemark[$\\dagger$] {\\it Chandra} observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] Start and end time of the observations. \\par\\noindent \\footnotemark[$\\S$] FAINT and VFAINT denote the imaging mode, and CC33\\_FAINT denotes the timing mode. \\par\\noindent \\footnotemark[$\\|$] Net exposure time. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\scriptsize \\caption{XMM-Newton observations of the quiescent emissions of the AXPs.}\\label{tab:axp_obs_xmm} \\begin{center} \\begin{tabular}{lllllllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & \\multicolumn{2}{l}{Observation Date (MJD)\\footnotemark[$\\ddagger$]} & \\multicolumn{3}{l}{Observation Mode\\footnotemark[$\\S$]} & \\multicolumn{3}{l}{Exposure Time (ks)\\footnotemark[$\\|$]} & \\multicolumn{3}{l}{Source/Background Radii} \\\\ & & Start & End & pn & MOS1 & MOS2 & pn & MOS1 & MOS2 & pn & MOS1 & MOS2 \\\\ \\hline 0100$-$721 & 0110000201 & 51834.626 & 51834.867 & E-Full & Full & Full & 20.81 & 14.62 & $g$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 0100$-$721 & 0018540101 & 52233.983 & 52234.303 & Full & Full & Full & 21.16 & 25.73 & $g$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\timeform{28''}$/$\\timeform{28''}$ & $\\timeform{28''}$/$\\timeform{28''}$ \\\\ 0142$+$614 & 0112781101 & 52663.920 & 52663.995 & Small & Fast-U & Fast-U & 4.18 & $t$ & $t$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 1708$-$400 & 0148690101 & 52879.906 & 52880.426 & Full & P-W3 & P-W3 & 26.88 & $p$ & $p$ & $\\timeform{20''}$/$\\timeform{20''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 1841$-$045 & 0013340101 & 52552.122 & 52552.192 & Large & Full & Full & 2.34 & 3.54 & 3.57 & $\\timeform{12''}$/$\\timeform{12''}$ & $\\timeform{12''}$/$\\timeform{12''}$ & $\\timeform{12''}$/$\\timeform{12''}$ \\\\ 1841$-$045 & 0013340201 & 52554.115 & 52554.193 & Large & Full & Full & 4.37 & 6.30 & 6.30 & $\\timeform{12''}$/$\\timeform{12''}$ & $\\timeform{12''}$/$\\timeform{12''}$ & $\\timeform{12''}$/$\\timeform{12''}$ \\\\ 2259$+$586 & 0038140101 & 52436.378 & 52436.986 & Small & Full & Full & 30.63 & $p$ & $p$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 2259$+$586 & 0155350301 & 52446.400 & 52446.759 & Small & P-W2 & Full & 17.65 & $p$ & $p$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 2259$+$586 & 0203550701 & 53579.965 & 53580.030 & Small & P-W2 & Fast-U & 2.66 & $p$ & $t$ & $\\timeform{32''}$/$\\timeform{32''}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ \\hline \\multicolumn{13}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the AXPs; CXOU\\,J010043.1$-$721134 (0100$-$721), 4U\\,0142$+$614 (0142$+$614), 1RXS\\,J170849.0$-$400910 (1708$-$400), 1E\\,1841$-$045 and 1E\\,2259$+$586. \\par\\noindent \\footnotemark[$\\dagger$] XMM-Newton observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] Start and end time of the observations. \\par\\noindent \\footnotemark[$\\S$] Observation mode for each instrument; extended full-window mode (E-Full), Full-window mode (Full), small-window mode (Small), fast-uncompressed mode (Fast-U), fast-timing mode (Fast-T), partial-w3 mode (P-W3), large-window mode (Large) and partial-w2 mode (P-W2). \\par\\noindent \\footnotemark[$\\|$] Net exposure time for each instrument. $g$ denotes that the source fell on a gap of the CCD chips, $t$ denotes observations in timing mode, and $p$ denotes that the data sets are affected by a photon pile-up. These data sets were not utilized. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\small \\caption{{\\it Chandra} observations of the quiescent emissions of the AXPs.}\\label{tab:axp_obs_cxo} \\begin{center} \\begin{tabular}{lllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & \\multicolumn{2}{l}{Observation Date (MJD)\\footnotemark[$\\ddagger$]} & Observation Mode\\footnotemark[$\\S$] & Exposure Time (ks)\\footnotemark[$\\|$] & Source/Background Radii \\\\ & & Start & End & & & \\\\ \\hline 0100$-$721 & 1881 & 52044.080 & 52045.261 & FAINT & 98.67 & $\\timeform{11''}$/$\\timeform{10''}$\\footnotemark[$\\#$] \\\\ 0100$-$721 & 4616 & 53031.791 & 53032.009 & VFAINT & 15.56 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 0100$-$721 & 4617 & 53032.189 & 53032.399 & VFAINT & 15.27 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 0100$-$721 & 4618 & 53033.904 & 53034.130 & VFAINT & 15.00 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 0100$-$721 & 4619 & 53042.806 & 53043.023 & VFAINT & 15.04 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 0100$-$721 & 4620 & 53089.185 & 53089.395 & VFAINT & 15.22 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 1647$-$455 & 6283 & 53512.860 & 53513.102 & FAINT & 18.81 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ 1647$-$455 & 5411 & 53539.673 & 53540.141 & FAINT & 38.47 & $\\timeform{2''}$/$\\timeform{2''}$ \\\\ \\hline \\multicolumn{7}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the AXPs; CXOU\\,J010043.1$-$721134 (0100$-$721) and CXOU\\,J164710.2$-$455216 (1647$-$455). \\par\\noindent \\footnotemark[$\\dagger$] {\\it Chandra} observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] Start and end time of observations. \\par\\noindent \\footnotemark[$\\S$] FAINT and VFAINT denote the imaging mode. \\par\\noindent \\footnotemark[$\\|$] Net exposure time. \\par\\noindent \\footnotemark[$\\#$] Since the source fell on an off-axis CCD chip, the source region was extracted from an elliptical region with major and minor axes of $\\timeform{11''}$ and $\\timeform{10''}$, respectively. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\small \\caption{{\\it Swift} observations of the quiescent emission and the bursts of AXP\\,CXOU\\,J164710.2$-$455216.}\\label{tab:axp_j1647_obs_swift} \\begin{center} \\begin{tabular}{lllllllll} \\hline\\hline SeqNum\\footnotemark[$*$] & \\multicolumn{4}{l}{Observation Date (MJD)\\footnotemark[$\\dagger$]} & \\multicolumn{2}{l}{Exposure Time (ks)\\footnotemark[$\\ddagger$]} & \\multicolumn{2}{l}{Source/Background Radii} \\\\ & Start (WT) & End (WT) & Start (PC) & End (PC) & WT & PC & WT & PC \\\\ \\hline 00230341000\\footnotemark[$\\S$] & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ \\\\ 00030806001 & 53999.604 & 53999.924 & 53999.604 & 53999.936 & 1.92 & 7.74 & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\timeform{30''}$/$\\timeform{30''}$ \\\\ 00030806002 & 54000.610 & 54000.619 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.77 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806003 & 54000.819 & 54001.212 & 54000.819 & 54001.073 & 4.91 & 1.84 & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\timeform{30''}$/$\\timeform{30''}$ \\\\ 00030806004 & 54004.276 & 54004.483 & 54004.343 & 54004.489 & 1.25 & 2.48 & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\timeform{30''}$/$\\timeform{30''}$ \\\\ 00030806006 & 54010.461 & 54010.716 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 1.98 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806007 & 54011.515 & 54011.594 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 2.03 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806008 & 54014.001 & 54014.077 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 2.16 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806009 & 54017.746 & 54017.954 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 3.52 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806010 & 54018.009 & 54018.094 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 2.83 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806011 & 54023.237 & 54023.517 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 5.62 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806012 & 54029.125 & 54029.342 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 5.52 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806013 & 54035.677 & 54035.825 & 54035.678 & 54035.774 & 2.82 & 0.42 & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\timeform{30''}$/$\\timeform{30''}$ \\\\ 00030806014 & 54119.174 & 54119.253 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 2.06 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ 00030806015 & 54122.050 & 54122.267 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 3.82 & $\\cdot\\cdot\\cdot$ & $\\timeform{36''}\\times\\timeform{18''}$/$\\timeform{36''}\\times\\timeform{18''}$\\footnotemark[$\\|$] & $\\cdot\\cdot\\cdot$ \\\\ \\hline \\multicolumn{9}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] {\\it Swift} sequence number. \\par\\noindent \\footnotemark[$\\dagger$] Start and end time of the observations for each mode (WT denotes window timing mode, and PC denotes photon counting mode). \\par\\noindent \\footnotemark[$\\ddagger$] Net exposure time for each observation mode. \\par\\noindent \\footnotemark[$\\S$] The observation of a burst. \\par\\noindent \\footnotemark[$\\|$] The source and background regions were extracted from a rectangle region. Two background regions are utilized near both sides of the source region. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{Spectral parameters of the quiescent emissions of the SGRs observed by XMM-Newton.}\\label{tab:sgr_spc_xmm} \\begin{center} \\begin{tabular}{lllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & $N_{\\mathrm{H}}$\\footnotemark[$\\ddagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\|$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\|$] & $F$\\footnotemark[$\\#$] & $\\chi^{2}$ (d.o.f.) \\\\ & & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 1627$-$41 & 0204500201 & 15.98$_{-7.60}^{+15.51}$ & 0.58$_{-0.25}^{+0.35}$ & $<$19.25 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\sim$\\,0.05 & 44 (42) \\\\ 1627$-$41 & 0204500301 & 7.53$_{-3.56}^{+6.19}$ & 0.85$_{-0.22}^{+0.29}$ & $<$0.62 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.08$\\pm0.06$ & 29 (65) \\\\ 1627$-$41 & 0202560101 & 9.00$_{-3.89}^{+6.71}$ & 0.94$_{-0.24}^{+0.31}$ & $<$0.42 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.06$\\pm0.04$ & 54 (47) \\\\ 1806$-$20 & 0148210101 & 5.18$_{-0.73}^{+0.92}$ & 0.84$_{-0.17}^{+0.23}$ & 1.64$_{-0.53}^{+1.04}$ & 2.62$_{-0.38}^{+0.97}$ & 0.28$_{-0.12}^{+0.10}$ & 11.05$\\pm1.92$ & 312 (295) \\\\ 1806$-$20 & 0148210401 & 5.87$_{-0.55}^{+0.63}$ & 0.85$_{-0.11}^{+0.12}$ & 1.97$_{-0.43}^{+0.65}$ & 3.39$_{-0.58}^{+1.2}$ & 0.20$\\pm0.07$ & 12.29$\\pm2.4$ & 506 (459) \\\\ 1806$-$20 & 0205350101 & 5.75$_{-0.19}^{+0.20}$ & 0.96$\\pm0.05$ & 2.06$_{-0.17}^{+0.20}$ & 3.19$_{-0.21}^{+0.28}$ & 0.32$\\pm0.04$ & 25.43$\\pm0.52$ & 2243 (2224) \\\\ 1806$-$20 & 0164561101 & 5.43$_{-0.35}^{+0.39}$ & 1.02$\\pm0.1$ & 1.94$_{-0.27}^{+0.37}$ & 3.92$_{-0.69}^{+1.46}$ & 0.23$\\pm0.08$ & 24.65$\\pm2.74$ & 737 (741) \\\\ 1806$-$20 & 0164561301 & 5.64$_{-0.43}^{+0.50}$ & 0.96$\\pm0.1$ & 1.98$_{-0.33}^{+0.46}$ & 3.67$_{-0.68}^{+1.54}$ & 0.22$\\pm0.08$ & 18.91$\\pm4.21$ & 654 (531) \\\\ 1806$-$20 & 0164561401 & 5.91$_{-0.31}^{+0.34}$ & 0.87$\\pm0.06$ & 2.03$_{-0.25}^{+0.31}$ & 3.27$_{-0.34}^{+0.48}$ & 0.22$\\pm0.04$ & 13.30$\\pm0.5$ & 1026 (1069) \\\\ 2013$+$34 & 0212481201 & 0.29$_{-0.13}^{+0.15}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.13$_{-0.02}^{+0.02}$ & $<$7.54 & 20.78$\\pm20.12$ & 57 (64) \\\\ 1819$-$16 & 0152834501 & $1.6_{-0.5}^{+0.7}$ & $1.9_{-0.2}^{+0.3}$ & $0.11_{-0.02}^{+0.03}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 1.3$\\pm0.1$ & 59 (74) \\\\ \\hline \\multicolumn{9}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the SGRs (2013$+$34 denotes a SGR candidate SGR/GRB\\,050925). \\par\\noindent \\footnotemark[$\\dagger$] XMM-Newton observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\#$] $F$ denotes a flux in the energy range 2-10 keV in units of $10^{-12}$ ergs cm$^{-2}$ s$^{-1}$ with 68 \\% confidence level errors. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{Spectral parameters of the quiescent emissions of the SGRs observed by {\\it Chandra}.}\\label{tab:sgr_spc_cxo} \\begin{center} \\begin{tabular}{lllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & $N_{\\mathrm{H}}$\\footnotemark[$\\ddagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\|$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\|$] & $F$\\footnotemark[$\\#$] & $\\chi^{2}$ (d.o.f.) \\\\ & & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 0526$-$66 & 747 & 0.20$_{-0.02}^{+0.03}$ & 0.36$\\pm0.03$ & 10.87$_{-1.26}^{+1.66}$ & 0.98$_{-0.14}^{+0.22}$ & 1.06$_{-0.37}^{+0.46}$ & 0.49$\\pm0.06$ & 180 (181) \\\\ 0526$-$66 & 1957 & 0.26$_{-0.04}^{+0.05}$ & 0.30$\\pm0.04$ & 15.08$_{-2.87}^{+4.93}$ & 0.70$_{-0.07}^{+0.11}$ & 2.24$_{-0.72}^{+0.85}$ & 0.39$\\pm0.06$ & 176 (185) \\\\ 1627$-$41 & 1981 & 8.47$_{-4.86}^{+6.41}$ & 0.89$_{-0.28}^{+0.57}$ & $<$0.67 & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $<$0.07 & 36 (34) \\\\ 1627$-$41 & 3877 & 17.34$_{-7.36}^{+9.89}$ & 0.42$_{-0.13}^{+0.2}$ & 3.56$_{-2.99}^{+33.39}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $<$0.06 & 22 (21) \\\\ \\hline \\multicolumn{9}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the SGRs. \\par\\noindent \\footnotemark[$\\dagger$] {\\it Chandra} observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\#$] $F$ denotes the flux in the 2-10\\,keV band in units of $10^{-12}$ ergs cm$^{-2}$ s$^{-1}$ with 68 \\% confidence level errors. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{Spectral parameters of a burst of SGR\\,2013$+$34 (GRB/SGR\\,050925) observed by {\\it Swift}.}\\label{tab:sgr_050925_spc_swift} \\begin{center} \\begin{tabular}{llllllll} \\hline\\hline SeqNum\\footnotemark[$*$] & $N_{\\mathrm{H}}$\\footnotemark[$\\dagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\ddagger$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\ddagger$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $F$\\footnotemark[$\\|$] & $\\chi^{2}$ (d.o.f.) \\\\ & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 00156838000 & $\\cdot\\cdot\\cdot$ & 6.6$_{-3.9}^{+5.6}$ & 3.1$_{-1.7}^{+3.7}$ & 20$_{-4}^{+15}$ & $\\gtrsim$0.2 & 0.81$\\pm0.28$ & 27 (25) \\\\ \\hline \\multicolumn{8}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] {\\it Swift} sequence number. \\par\\noindent \\footnotemark[$\\dagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\ddagger$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $F$ denotes a flux in the energy range 15-150 keV in units of $10^{-6}$ ergs cm$^{-2}$ s$^{-1}$ with 68 \\% confidence level errors. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\small \\caption{Spectral parameters of the quiescent emissions of the AXPs observed by XMM-Newton.}\\label{tab:axp_spc_xmm} \\begin{center} \\begin{tabular}{lllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & $N_{\\mathrm{H}}$\\footnotemark[$\\ddagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\|$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\|$] & $F$\\footnotemark[$\\#$] & $\\chi^{2}$ (d.o.f.) \\\\ & & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 0100$-$721 & 0110000201 & $\\lesssim0.06$ & 0.39$\\pm0.02$ & 6.94$_{-0.67}^{+1.09}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.09$\\pm0.01$ & 264 (275) \\\\ 0100$-$721 & 0018540101 & $\\lesssim0.15$ & 0.29$\\pm0.06$ & 11.64$_{-3.35}^{+7.47}$ & 0.64$_{-0.11}^{+0.25}$ & 1.86$_{-1.08}^{+1.17}$ & 0.13$\\pm0.09$ & 97 (129) \\\\ 0142$+$614 & 0112781101 & 0.53$\\pm0.01$ & 0.36$\\pm0.01$ & 9.38$_{-0.31}^{+0.34}$ & 0.82$_{-0.02}^{+0.03}$ & 0.89$_{-0.08}^{+0.09}$ & 57.26$\\pm0.61$ & 1086 (819) \\\\ 1708$-$400 & 0148690101 & 0.95$\\pm0.01$ & 0.48$\\pm0.01$ & 4.46$_{-0.10}^{+0.11}$ & 1.49$\\pm0.04$ & 0.29$\\pm0.02$ & 27.42$_{-0.13}^{+0.08}$ & 1566 (1232) \\\\ 1841$-$045 & 0013340101 & 1.86$_{-0.13}^{+0.14}$ & 0.52$\\pm0.03$ & 3.79$_{-0.46}^{+0.57}$ & 1.99$_{-0.20}^{+0.25}$ & 0.21$_{-0.04}^{+0.05}$ & 17.39$\\pm0.83$ & 408 (391) \\\\ 1841$-$045 & 0013340201 & 1.90$\\pm0.11$ & 0.51$\\pm0.02$ & 3.97$_{-0.40}^{+0.48}$ & 1.81$_{-0.12}^{+0.14}$ & 0.25$_{-0.03}^{+0.04}$ & 17.16$\\pm0.51$ & 583 (641) \\\\ 2259$+$586 & 0038140101 & $0.59\\pm0.01$ & $0.353_{-0.004}^{+0.005}$ & $4.88_{-0.13}^{+0.15}$ & $0.75\\pm0.02$ & $0.50_{-0.04}^{+0.05}$ & $12.34\\pm0.07$ & 1167 (888) \\\\ 2259$+$586 & 0155350301 & $0.54\\pm0.01$ & $0.390\\pm0.006$ & $5.01_{-0.14}^{+0.15}$ & $0.85\\pm0.02$ & $0.67_{-0.04}^{+0.05}$ & $33.01\\pm0.21$ & 1531 (1053) \\\\ 2259$+$586 & 0203550701 & $0.59\\pm0.03$ & $0.356_{-0.015}^{+0.013}$ & $4.97_{-0.39}^{+0.47}$ & $0.82_{-0.07}^{+0.08}$ & $0.43_{-0.10}^{+0.13}$ & $13.77_{-0.44}^{+0.44}$ & 457 (458) \\\\ \\hline \\multicolumn{9}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the AXPs; CXOU\\,J010043.1$-$721134 (0100$-$721), 4U\\,0142$+$614 (0142$+$614), 1RXS\\,J170849.0$-$400910 (1708$-$400), 1E\\,1841$-$045 and 1E\\,2259$+$586. \\par\\noindent \\footnotemark[$\\dagger$] XMM-Newton observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\#$] $F$ denotes a flux in the energy range 2-10 keV in units of $10^{-12}$ ergs cm$^{-2}$ s$^{-1}$ with 68 \\% confidence level errors. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{Spectral parameters of the quiescent emissions of the AXPs observed by {\\it Chandra}.}\\label{tab:axp_spc_cxo} \\begin{center} \\begin{tabular}{lllllllll} \\hline\\hline Object\\footnotemark[$*$] & ObsID\\footnotemark[$\\dagger$] & $N_{\\mathrm{H}}$\\footnotemark[$\\ddagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\|$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\|$] & $F$\\footnotemark[$\\#$] & $\\chi^{2}$ (d.o.f.) \\\\ & & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 0100$-$721 & 1881 & 0.06$_{-0.05}^{+0.06}$ & 0.33$_{-0.04}^{+0.04}$ & 9.65$_{-1.74}^{+2.98}$ & 0.65$_{-0.11}^{+0.23}$ & 1.42$_{-0.85}^{+1.25}$ & 0.12$_{-0.05}^{+0.05}$ & 172 (150) \\\\ 0100$-$721 & 4616 & $<0.04$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.39$\\pm0.02$ & 6.84$_{-0.56}^{+1.06}$ & 0.09$\\pm0.01$ & 63 (68) \\\\ 0100$-$721 & 4617 & $<0.41$ & 0.18$_{-0.06}^{+0.25}$ & 21.90$_{-20.07}^{+173.99}$ & 0.44$_{-0.05}^{+21.23}$ & 5.53$_{-5.52}^{+1.24}$ & 0.12$\\pm0.04$ & 66 (64) \\\\ 0100$-$721 & 4618 & $<0.18$ & 0.28$\\pm0.1$ & 10.88$_{-3.73}^{+15.89}$ & 0.51$_{-0.09}^{+1.04}$ & 3.02$_{-2.85}^{+2.67}$ & 0.10$\\pm0.09$ & 50 (65) \\\\ 0100$-$721 & 4619 & $<0.04$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.41$_{-0.2}^{+0.02}$ & 6.40$_{-0.51}^{+0.88}$ & 0.11$\\pm0.01$ & 47 (67) \\\\ 0100$-$721 & 4620 & $<0.03$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.40$_{-0.02}^{+0.01}$ & 6.67$_{-0.41}^{+0.79}$ & 0.10$\\pm0.01$ & 70 (68) \\\\ 1647$-$455 & 6283 & 2.54$_{-0.69}^{+0.81}$ & 0.49$\\pm0.06$ & 0.52$_{-0.18}^{+0.31}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.15$\\pm0.04$ & 23 (21) \\\\ 1647$-$455 & 5411 & 1.44$_{-0.28}^{+0.32}$ & 0.58$\\pm0.05$ & 0.26$_{-0.05}^{+0.07}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 0.13$\\pm0.01$ & 54 (44) \\\\ \\hline \\multicolumn{9}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] Object name of the AXPs; CXOU\\,J010043.1$-$721134 (0100$-$721) and CXOU\\,J164710.2$-$455216 (1647$-$455). \\par\\noindent \\footnotemark[$\\dagger$] {\\it Chandra} observation ID. \\par\\noindent \\footnotemark[$\\ddagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\#$] $F$ denotes a flux in the energy range 2-10 keV in units of $10^{-12}$ ergs cm$^{-2}$ s$^{-1}$ with 68 \\% confidence level errors. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{Spectral parameters of the quiescent emission and the bursts of AXP\\,CXOU\\,J164710.2$-$455216 observed by {\\it Swift}.}\\label{tab:axp_j1647_spc_swift} \\begin{center} \\begin{tabular}{llllllll} \\hline\\hline SeqNum\\footnotemark[$*$] & $N_{\\mathrm{H}}$\\footnotemark[$\\dagger$] & $kT_{\\mathrm{LT}}$\\footnotemark[$\\ddagger$] & $R_{\\mathrm{LT}}$\\footnotemark[$\\S$] & $kT_{\\mathrm{HT}}$\\footnotemark[$\\ddagger$] & $R_{\\mathrm{HT}}$\\footnotemark[$\\S$] & $F$\\footnotemark[$\\|$] & $\\chi^{2}$ (d.o.f.) \\\\ & (10$^{22}$ cm$^{-2}$) & (keV) & (km) & (keV) & (km) & & \\\\ \\hline 00230341000\\footnotemark[$\\#$] & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 8.92$_{-1.62}^{+1.34}$ & 1.21$_{-0.35}^{+0.59}$ & 0.35$\\pm0.12$ & 8 (12) \\\\ 00030806001 & 1.72$_{-0.20}^{+0.31}$ & 0.63$_{-0.12}^{+0.07}$ & 2.40$_{-0.44}^{+0.94}$ & 2.08$_{-0.93}^{+38.74}$ & 0.14$_{-0.13}^{+0.36}$ & 27.94$\\pm10.25$ & 253 (242) \\\\ 00030806002 & 1.76$_{-0.55}^{+0.70}$ & 0.73$_{-0.08}^{+0.08}$ & 1.79$_{-0.44}^{+0.7}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 20.01$\\pm4.36$ & 16 (17) \\\\ 00030806003 & 1.95$_{-0.53}^{+0.86}$ & 0.42$_{-0.15}^{+0.26}$ & 3.03$_{-1.9}^{+9.21}$ & 0.81$_{-0.09}^{+0.35}$ & 1.04$_{-0.81}^{+0.43}$ & 13.98$\\pm6.28$ & 101 (92) \\\\ 00030806004 & 1.79$_{-0.7}^{+0.91}$ & 0.51$_{-0.21}^{+0.28}$ & 2.20$_{-1.66}^{+4.56}$ & 0.93$_{-0.38}^{+3.06}$ & 0.68$_{-0.67}^{+0.5}$ & 13.79$\\pm12.1$ & 91 (74) \\\\ 00030806006 & 0.99$_{-0.32}^{+0.37}$ & 0.75$\\pm0.06$ & 1.24$_{-0.23}^{+0.31}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 11.99$\\pm1.2$ & 31 (33) \\\\ 00030806007 & 1.65$_{-0.39}^{+0.49}$ & 0.67$_{-0.06}^{+0.07}$ & 1.52$_{-0.33}^{+0.48}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 9.85$\\pm1.34$ & 27 (29) \\\\ 00030806008 & 1.78$_{-0.42}^{+0.55}$ & 0.60$\\pm0.05$ & 2.02$_{-0.46}^{+0.69}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 9.18$\\pm1.33$ & 73 (64) \\\\ 00030806009 & 2.09$_{-0.46}^{+0.57}$ & 0.62$\\pm0.06$ & 1.69$_{-0.39}^{+0.58}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 6.86$\\pm1.38$ & 64 (55) \\\\ 00030806010 & 1.51$_{-0.32}^{+0.4}$ & 0.71$\\pm0.06$ & 1.31$_{-0.25}^{+0.34}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 9.48$\\pm0.81$ & 47 (59) \\\\ 00030806011 & 1.35$_{-0.24}^{+0.28}$ & 0.67$\\pm0.04$ & 1.38$_{-0.20}^{+0.26}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 7.93$\\pm0.61$ & 90 (93) \\\\ 00030806012 & 1.30$_{-0.24}^{+0.29}$ & 0.67$\\pm0.05$ & 1.22$_{-0.20}^{+0.25}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 6.60$\\pm0.45$ & 66 (75) \\\\ 00030806013 & 1.79$_{-0.40}^{+0.49}$ & 0.63$\\pm0.06$ & 1.57$_{-0.36}^{+0.54}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 7.00$\\pm0.97$ & 63 (70) \\\\ 00030806014 & 1.57$_{-0.51}^{+0.72}$ & 0.63$\\pm0.09$ & 1.26$_{-0.38}^{+0.68}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 4.64$\\pm1.43$ & 22 (31) \\\\ 00030806015 & 1.80$_{-0.42}^{+0.53}$ & 0.61$_{-0.06}^{+0.07}$ & 1.38$_{-0.34}^{+0.53}$ & $\\cdot\\cdot\\cdot$ & $\\cdot\\cdot\\cdot$ & 4.77$\\pm0.84$ & 49 (47) \\\\ \\hline \\multicolumn{8}{@{}l@{}}{\\hbox to 0pt{\\parbox{180mm}{\\footnotesize \\footnotemark[$*$] {\\it Swift} sequence number. \\par\\noindent \\footnotemark[$\\dagger$] $N_{\\mathrm{H}}$ denotes the column density with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\ddagger$] $kT_{\\mathrm{LT}}$ and $kT_{\\mathrm{HT}}$ denote the blackbody temperatures with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\S$] $R_{\\mathrm{LT}}$ and $R_{\\mathrm{HT}}$ denote the emission radii with 90 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\|$] $F$ denotes fluxes in the energy ranges 15-150 keV in units of $10^{-6}$ ergs cm$^{-2}$ s$^{-1}$ for the burst observation of 00230341000 and 2-10 keV in units of $10^{-12}$ ergs cm$^{-2}$ s$^{-1}$ for other observations with 68 \\% confidence level errors. \\par\\noindent \\footnotemark[$\\#$] Results for the burst. }\\hss}} \\end{tabular} \\end{center} \\end{table*} \\clearpage %%% % See the manual for the detail. %%%" }, "0710/0710.1778_arXiv.txt": { "abstract": "Using Magellan/IMACS images covering a 1.2 x 1.2 sq. degree FOV with seeing of 0.4\"-0.6\", we have applied convolution techniques to analyse the light distribution of 364 confirmed globular cluster in the field of NGC 5128 and to obtain their structural parameters. Combining these parameters with existing Washington photometry from Harris et al. (2004), we are able to examine the size difference between metal-poor (blue) and metal-rich (red) globular clusters. For the first time, this can be addressed on a sample of confirmed clusters that extends to galactocentric distances about 8 times the effective radius, R$_{eff}$, of the galaxy. Within 1 R$_{eff}$, red clusters are about $30\\%$ smaller on average than blue clusters, in agreement with the vast majority of extragalactic globular cluster systems studied. As the galactocentric distance increases, however, this difference becomes negligible. Thus, our results indicate that the difference in the clusters' effective radii, r$_e$, could be explained purely by projection effects, with red clusters being more centrally concentrated than blue ones and an intrinsic r$_e$--R$_{gc}$ dependence, like the one observed for the Galaxy. ", "introduction": "\\label{sec:intro} Since sizes and structural parameters of globular clusters (GCs) in different GC systems (GCSs) have first been obtained, it has become clear that some of these properties correlate with global properties of their host galaxies \\citep[see for example][]{jordan05,brodie06}. The existence of the so called fundamental plane relation for an increasing number of studied GCSs seems to confirm that GCs populate a narrow region in this parameter space \\citep{djorgovski95,mclaughlin00,mclaughlin05,barmby07}. However, there are puzzling trends that are still awaiting confirmation and need to be addressed using larger samples of GCs. It is necessary to study structural parameters of GCs and GC-like objects in different environments before definitive statements can be made regarding their formation. Among the structural parameters that can be studied, the effective (or half-light) radius is of particular importance. Models have shown that this quantity remains fairly constant throughout the entire GC lifetime \\citep{spitzer72,aarseth98}, making it a good indicator of proto-GC sizes that are still observable today. A decade ago, HST observations unveiled a systematic size difference between red and blue GCs \\citep{kundu98}. Since then, multiple studies have found that the blue GCs are between $17\\%-30\\%$ larger than their metal-rich counterparts in both spirals and early--type galaxies \\citep{kundu99,puzia99,larsen01,larsen_fb01,kundu01,barmby02,jordan05}. However, most of these studies have made use of HST observations and examine only the innermost regions of the galaxy or small fields in regions at galactocentric distances greater than the galaxy's effective radius. According to \\cite{larsen03}, the systematic size difference between red and blue GCs is caused merely by a projection effect. Since red (metal-rich) GCs are found to be more centrally concentrated than blue (metal-poor ones) in early type galaxies \\citep[][among others]{cote01,dirsch03,woodley05}, the red GCs will appear to lie, on average, at a smaller galactocentric distance. The red clusters will on average be smaller than the blue clusters assuming that both types shares the same relation between the GC size and galactocentric distance. The relation r $\\sim \\sqrt{\\rm{R}_{gc}}$ was first found in the Milky Way by \\cite{vandenbergh91}. In this scenario, the difference between the cluster sizes should be most apparent at small galactocentric distance and should decrease strongly beyond 1 galaxy effective radius \\citep{larsen03}. Alternatively, \\cite{jordan04} suggests that this effect could be explained by an intrinsic difference between metal-rich and metal-poor GCs. Assuming half-mass radii that are independent of metallicity, effects of mass segregation combined with a metallicity-dependent stellar lifetime should lead to different sizes between the blue and red clusters. The brightest stars would be more massive and more centrally concentrated for the metal-rich GCs. This scenario should have little to no dependence on a cluster's distance from the center of its parent galaxy. In a recent study, \\cite{spitler06} analysed the GCS of NGC 4594 (Sombrero, at a distance of $ \\sim 9$ Mpc) using a six-image mosaic from HST/ACS. They confirm that within the inner 2 arcmin (2.2 R$_{eff}$), the metal-rich GCs are, on average, $17\\%$ smaller than the metal-poor clusters. However, the size difference becomes negligible at $\\sim 3$ arcmin, corresponding to $\\sim 3.4$ R$_{eff}$, where R$_{eff} =0.89$ arcmin \\citep{baggett98}. To further understand the sizes of red and blue clusters, we need a homogeneous survey of a GCS with the ability to eliminate contaminating sources, high resolution to measure structural parameters, and over a large range in galactocentric distance. NGC 5128 is the nearest giant elliptical galaxy, at a distance of 3.8 Mpc \\citep{mclaughlin07}. Its GCs are thus easily resolvable with sub-arcsecond seeing \\citep{harris06}. In this paper we present effective radius results for 337 GCs from the \\cite{woodley07} catalog that are confirmed GCs by either radial velocity measurement from various studies \\citep[see the references in ][]{woodley07} or are resolved by HST/ACS images \\citep{harris06}. We also present the effective radii of 27 GCs newly confirmed through radial velocity measurements using the Baade 6.5-m telescope with the instrument LDSS-2 (data in preparation for publication). This list represents a clean sample of confirmed clusters. All of these also have ellipticities less than 0.4 and effective radii less than 8 pc, both of which are consistent with {\\it normal} GC properties in NGC 5128. We find that only an additional $2.4\\%$ of GCs from the \\cite{woodley07} catalog have effective radii greater than the 8 pc boundary we have imposed here (to be discussed in detail in G\\'omez \\& Woodley, 2008, in preparation). Those few GCs are not considered here as our purpose is to establish the effective radius trends within the bulk of the GC population. ", "conclusions": "\\label{sec:conclusions} Using a contaminant-free sample of 364 GCs in NGC 5128, confirmed with radial velocity measurements or by resolved HST images, we have measured effective radii using ISHAPE. Our results indicate that the blue or metal-poor clusters do not show any significant r$_{e}$--R$_{gc}$ relation. However, the red or metal-rich GCs do show a steep relation in which red clusters within 1 R$_{eff}$ of the galaxy's light are $30\\%$ smaller than the blue clusters. Beyond this distance there is no indication for a size difference between the two metallicity populations. This finding in NGC 5128, not previously seen in any other early-type galaxy, supports the more tentative findings of the Sombrero galaxy's GCS \\citep{spitler06}. Both studies support the idea that the size differences are most likely caused by projection effects \\citep{larsen03} and not by intrinsic physical differences between the two subgroups. Acknowledgements: M.G. and K.A.W. thank Dean McLaughlin for use of HST structural parameters in advance of publication. M.G. thanks the Dept. of Physics and Astronomy at McMaster University and especially Bill and Gretchen Harris for their hospitality. K.A.W. thanks NSERC and Bill Harris for financial support, and also the Depto. de F{\\'i}sica at the Universidad de Concepci{\\'o}n, especially Doug Geisler, for their hospitality. We thank the anonymous referee for her/his valuable suggestions and comments. \\clearpage" }, "0710/0710.3850_arXiv.txt": { "abstract": "The first map of interstellar acetylene (C$_2$H$_2$) has been obtained with the infrared spectrograph onboard the {\\it Spitzer Space Telescope}. A spectral line map of the $\\nu_5$ vibration-rotation band at 13.7 $\\mu$m carried out toward the star-forming region Cepheus A East, shows that the C$_2$H$_2$ emission peaks in a few localized clumps where gas-phase CO$_2$ emission was previously detected with {\\it Spitzer}. The distribution of excitation temperatures derived from fits to the C$_2$H$_2$ line profiles ranges from 50 to 200 K, a range consistent with that derived for gaseous CO$_2$ suggesting that both molecules probe the same warm gas component. The C$_2$H$_2$ molecules are excited via radiative pumping by 13.7 $\\mu$m continuum photons emanating from the HW2 protostellar region. We derive column densities ranging from a few $\\times$ 10$^{13}$ to $\\sim$ 7 $\\times$ 10$^{14}$ cm$^{-2}$, corresponding to C$_2$H$_2$ abundances of 1 $\\times$ 10$^{-9}$ to 4 $\\times$ 10$^{-8}$ with respect to H$_2$. The spatial distribution of the C$_2$H$_2$ emission along with a roughly constant $N$(C$_2$H$_2$)/$N$(CO$_2$) strongly suggest an association with shock activity, most likely the result of the sputtering of acetylene in icy grain mantles. ", "introduction": "Acetylene constitutes a key ingredient in the production of large complex hydrocarbon molecules in the dense interstellar medium (Herbst 1995). Because acetylene has no permanent dipole moment, it lacks a rotational spectrum that could be observed at radio wavelengths; observations of interstellar acetylene have therefore been limited to mid-infrared studies of rovibrational bands, carried out from ground-(e.g., Evans, Lacy \\& Carr 1991; Carr et al. 1995) and space-based (e.g., Lahuis \\& van Dishoeck 2000; Boonman et al. 2003; Lahuis et al. 2007) observatories. Acetylene has been detected in the gas-phase - either in absorption (e.g., Carr et al. 1995; Lahuis \\& van Dishoeck 2000) or in emission (e.g. Boonman et al. 2003, this paper) - mostly toward young stellar objects. C$_2$H$_2$ can be used as a tracer of warm (100 K to 1000 K) molecular gas along often complicated sightlines. C$_2$H$_2$ abundance estimates, which were sometimes a few orders of magnitude higher than the predictions of cold gas-phase steady-state chemical models, have led to a better understanding of the role that warm gas-phase chemistry (e.g., Doty et al. 2002; Rodgers \\& Charnley 2001) and/or grain mantle processing (e.g., Ruffle \\& Herbst 2000) can play in star-forming regions, both locally and in extra-galactic objects (Lahuis et al. 2007). In this Letter, we present the first detection of the $\\nu_5$ band of acetylene (C$_2$H$_2$) at 13.7 $\\mu$m toward the star forming region Cepheus A East using the Infrared Spectrograph (IRS) onboard the {\\it Spitzer Space Telescope}. This is the first map of C$_2$H$_2$ obtained toward any interstellar gas cloud. Section 2 describes the observations and data analysis. Sections 3 and 4 compare the spatial distribution of C$_2$H$_2$ to those of gaseous CO$_2$ and H$_2$ $S$(2), and discuss the C$_2$H$_2$-emitting gas in the context of shock chemistry and local outflow activity. The presence of C$_2$H$_2$ on interstellar dust grains will also be discussed in the context of cometary ices composition. ", "conclusions": "Acetylene was previously observed toward low-to-high mass star-forming regions either in absorption (e.g., Lahuis \\& van Dishoeck 2000) or in emission (e.g., Boonman et al. 2003) using the {\\it Infrared Space Observatory} ({\\it ISO}). Excitation temperatures ranging from $\\sim$ 10 to 900 K and abundances with respect to H$_2$ ranging from a few $\\times$ 10$^{-8}$ to a few $\\times$ 10$^{-7}$ were derived. Steady-state models of gas-phase chemistry in cold (10-50K) dense (10$^3$-10$^5$ cm$^{-3}$) molecular clouds predict abundances for acetylene between a few $\\times$ 10$^{-10}$ and 1$\\times$ 10$^{-8}$ with respect to H$_2$ depending on the role that neutral-neutral destruction reactions may play (Bettens, Lee \\& Herbst 1995; Lee et al. 1996). Similar abundances are predicted by models of gas-grain chemistry in quiescent clouds with the highest values obtained only after 10$^6$ years (Ruffle \\& Herbst 2000). While such models could account for the observed abundances in the cold gas, they were unable to reproduce the enhancements observed toward much warmer gas components in those objects. Mechanisms such as C$_2$H$_2$ ice sublimation from grain mantles and/or C$_2$H$_2$ enhancements via warm gas-phase chemistry were then invoked (e.g., Carr et al. 1995; Doty et al. 2002; Boonman et al. 2003). For the NE outflow region in Cepheus A East, we derive C$_2$H$_2$/H$_2$ abundance ratios in the range 1 $\\times$ 10$^{-9}$ to 4 $\\times$ 10$^{-8}$, for an assumed H$_2$ column density of 1.5 $\\times$ 10$^{22}$ cm$^{-3}$ (G\\'omez et al. 1999). These values are averages along the observed sight-lines. The highest abundances are localized around and to the south of, the NE position as well as around the positions of the HW5/6 sources (see Fig.~2). This is precisely where the interaction between the {\\it northeast} outflow and the ambient molecular clouds occurs, and it is in these regions that we observe strong H$_2$ $S$(2) emission, a tracer of warm shocked gas. Thus the spatial variation of the C$_2$H$_2$ abundance again strongly suggests an association with shock activity, perhaps as a result of (1) production in the gas phase via high temperature reactions, or (2) grain mantle sputtering. Models for chemistry in hot cores (Rodgers \\& Charnley 2001; Doty et al. 2002) indicate that enhanced abundances of C$_2$H$_2$ ($\\sim$ few $\\times$ 10$^{-8}$) are expected in warm regions with $T \\ge$ 200 K. The good correlations between H$_2$ $S$(2), gaseous CO$_2$ and C$_2$H$_2$ indicate that such high temperatures were reached in the warm gas at the passage of the non-dissociative shock. However, the chemical pathways leading to the production of C$_2$H$_2$ are slow, and enhanced abundances only occur after $\\sim$ 10$^4$ years, a time scale much greater than that expected for shock heating of the gas ($\\sim$ 300 years; e.g. Kaufman \\& Neufeld 1996). Hence, enhanced production of C$_2$H$_2$ by high temperature gas-phase chemistry is unlikely to be predominant in the observed region. Thus the correlations shown in Fig.~3 argue in favor of grain mantle sputtering over gas-phase production as the origin of the C$_2$H$_2$ in Cepheus A East. This is the same production mechanism that we favored for gaseous CO$_2$ (Sonnentrucker et al. 2006). While models for the production of C$_2$H$_2$ in shocks are not available to our knowledge, our results further suggest that both C$_2$H$_2$ and CO$_2$ are released into the gas phase under very similar physical conditions. The gaseous $\\rm C_2H_2/CO_2$ ratio is roughly constant, with a mean value of 0.08 for all sight-lines where we detected acetylene at the 1.5 $\\sigma$ level. If this value reflects the composition of the grain mantle, and given a $N$(CO$_2$)$_{ice}$/$N$(H$_2$O)$_{ice}$ ratio in this source of 0.22 (Sonnentrucker et al. 2007, ApJ in press), then the required $N$(C$_2$H$_2$)$_{ice}$/$N$(H$_2$O)$_{ice}$ ratio is 0.02. This value is at least a factor 4 larger than those derived toward other star-forming regions (0.1-0.5\\%, Evans et al. 1991; Lahuis \\& van Dishoeck 2000) and those predicted by theoretical models (0.1-0.5\\%, Hasegawa \\& Herbst 1993; Ruffle \\& Herbst 2000), and at least a factor 2 larger than the gaseous C$_2$H$_2$/H$_2$O ratios obtained in observations of cometary comae (0.1-0.9\\%, Brooke et al. 1996; Weaver et al. 1999). We speculate that these discrepancies might result from (1) the destruction of CO$_2$ by reaction with atomic hydrogen in shocks faster than $\\sim$ 30 km s$^{-1}$ (predicted by Charnley \\& Kaufman, 2000) and/or (2) a greater efficiency for sputtering of C$_2$H$_2$ in slow shocks.\\footnote{Although both these effects would be strongly dependent upon the shock velocity, the relative constancy of the $\\rm C_2H_2/CO_2$ would not necessarily require any fine tuning of the shock velocity. In reality, any sight-line typically samples an ensemble of shocks with a {\\it range} of shock velocities, and the constancy of the $\\rm C_2H_2/CO_2$ would simply indicate that the admixture of shock velocities varies little from one sight-line to another (e.g. Neufeld et al.\\ 2006).} In either case, the gaseous C$_2$H$_2$/CO$_2$ ratios we observed may exceed the solid C$_2$H$_2$/CO$_2$ ratio, and the $N$(C$_2$H$_2$)$_{ice}$/$N$(H$_2$O)$_{ice}$ ratio could be less than 0.02. Unfortunately, direct measurements of acetylene ice are not possible, the weak features expected from solid C$_2$H$_2$ being blended with much stronger features of CO and H$_2$O (Boudin et al. 1998). However, further observations of gaseous C$_2$H$_2$ at a higher signal-to-noise ratio would be very valuable as a probe of any variations in the $\\rm C_2H_2/CO_2$ ratio which might provide important clues to the shock physics." }, "0710/0710.1064_arXiv.txt": { "abstract": "In this paper we report on the gas-phase abundance of singly-ionized magnesium (Mg II) in 44 lines of sight, using data from the {\\it Hubble Space Telescope} ({\\it HST}). We measure Mg II column densities by analyzing medium- and high-resolution archival STIS spectra of the 1240 \\AA{} doublet of Mg II. We find that Mg II depletion is correlated with many line of sight parameters (e.g.~$\\fHmol$, $E_{B-V}$, $\\ebvdist$, $A_V$, and $\\avdist$) in addition to the well-known correlation with $\\nHavg$. These parameters should be more directly related to dust content and thus have more physical significance with regard to the depletion of elements such as magnesium. We examine the significance of these additional correlations as compared to the known correlation between Mg II depletion and $\\nHavg$. While none of the correlations are better predictors of Mg II depletion than $\\nHavg$, some are statistically significant even assuming fixed $\\nHavg$. We discuss the ranges over which these correlations are valid, their strength at fixed $\\nHavg$, and physical interpretations. ", "introduction": "\\label{s:MgII_intro} Magnesium is both a relatively abundant element in the Galaxy and an important component in most interstellar dust models. Mg I has an ionization potential of only 7.65 eV, while Mg II has an ionization potential of 15.04 eV. In H I regions, the dominant form of gas-phase interstellar magnesium should be Mg II. In H II regions, magnesium should be found primarily in the form of Mg III. Gas-phase Mg I is rare, even in $\\Hmol$ regions. As is the case for other elements such as silicon and iron, the average gas-phase abundance of magnesium is much smaller than the assumed overall cosmic abundance of magnesium, implying that the majority of interstellar magnesium is tied up in dust. Therefore, variations in the gas-phase magnesium abundance are only capable of having minor effects on grain composition. Nevertheless, observed variations in the gas-phase magnesium abundance may still shed light on the physical conditions of interstellar clouds. Two major {\\it Copernicus} surveys that included measurements of Mg II abundances and depletions were \\citet{Murray} and \\citet{Jenkins1986}. Both of these studies confirmed that magnesium depletion increases with increased average hydrogen volume density, $\\nHavg=\\NHtot/r$ (where $r$ is the line-of-sight pathlength), in the line of sight. \\citet{Jenkins1986} also found several correlations between the depletions of magnesium and other elements, which they cited as secondary to the correlation between those depletions and $\\nHavg$. In both studies, the depletion of magnesium was not strongly correlated to other line of sight parameters, such as $\\ebv$ and the magnitude of the 2175 \\AA{} extinction ``bump''. We also note that there is a systematic difference in the absolute values of the abundances and depletions between these studies and more recent studies that is fairly substantial. This is due to a difference in the assumed $f$-values of the 1240 \\AA{} doublet of magnesium. We comment on our chosen $f$-values in \\S \\ref{ss:MgII_lines}. More recently, \\citet{Cartledge2006} also examined the abundances and depletions of Mg II and several other elements in the interstellar medium using STIS data. (Hereafter, this paper will be referred to as CLMS, for the initials of the authors.) Similar to the {\\it Copernicus} studies, CLMS concluded that the line of sight parameter with the clearest connection to elemental depletions is the average hydrogen volume density, $\\nHavg$. CLMS also explored potential correlations between depletions and other line of sight parameters such as the molecular fraction of hydrogen, $\\fHmol$; selective extinction, $\\ebv$; and selective extinction divided by line-of-sight pathlength, $\\ebvdist$. CLMS concluded that $\\nHavg$ was the parameter that best identified warm vs.~cold clouds, and that no other parameters produced correlations with magnesium depletion that were both as strong and with as little scatter. We began this study before CLMS was published. In spite of the similarities between that study and this one, we have proceeded with this study to provide an independent analysis of Mg II, and also because we have analyzed potential trends with respect to parameters not analyzed in CLMS, e.g.~$\\av$ and $\\rv$. The Mg II column densities of 11 out of the 44 lines of sight in our sample have also been analyzed previously by CLMS. We still report our results in this paper for two main reasons: (1) these lines of sight provide a basis of comparison for the methods of this paper and those of CLMS and (2) these lines of sight can be analyzed with respect to the aforementioned line of sight parameters not analyzed in CLMS. A comparison of the column density measurements for these common lines of sight is found in \\S \\ref{ss:MgII_coldensities}. In \\S \\ref{s:MgII_obsdata} we discuss our observations and data reduction, including comments on the 1240 \\AA{} doublet and our derivation of column densities and abundances. In \\S \\ref{s:results} we discuss our results, including observed correlations and a review of the Galactic abundance of magnesium. In \\S \\ref{s:MgII_summary} we summarize our findings. ", "conclusions": "\\label{s:results} \\subsection{Correlations} \\label{ss:correlations} Our results for the column densities of Mg II are given in Table \\ref{MgII_coldensities}. The major correlation found by both \\citet{Jenkins1986} and CLMS is that the depletion of magnesium and many other elements increases with overall line of sight density, $\\nHavg$. Both interpreted their models in light of the model by \\citet{Spitzer1985}. The \\citeauthor{Spitzer1985} model contends that the ISM contains two distinct varieties of cloud (warm and cold), each with a distinct depletion level. In this model, lines of sight with average densities of a few-tenths of a particle per cm$^{-3}$ are largely sampling the warm ISM, while lines of sight with average densities of a few particles per cm$^{-3}$ are largely sampling the cold ISM. This explains the observed plateaus of depletion at both low and high densities, with a transition near average densities of $\\sim1$ cm$^{-3}$, where lines of sight are sampling both types of clouds. This model, however, does not take into account the transition to a much different regime of cloud chemistry expected for translucent clouds \\citep[for a recent review, see][]{SnowMcCall}. \\citet{Jenkins1986} also found evidence of lesser correlations between the depletion of various elements and other parameters of reddening and extinction, but concluded that these correlations were secondary to the correlation with $\\nHavg$. CLMS examined the depletion of Mg II compared to $\\fHmol$, $\\ebv$, and $\\ebvdist$. CLMS concluded that only $\\ebvdist$ provided a correlation with magnesium depletion that was as at least roughly as significant as the correlation with $\\nHavg$, but with increased scatter. However, in our recent work \\citep{JensenFeII}, we explored the possibility of correlations between iron depletion and various other line of sight parameters. We found very similar correlations between iron depletion and measures of dust density ($\\ebvdist$ and $\\avdist$) and the molecular fraction of hydrogen ($\\fHmol$). We also examined iron depletion as a function of $\\rv$, the ratio of total visual extinction to selective extinction, though a conclusive trend did not present itself. Using many of the same lines of sight in this paper as in \\citet{JensenFeII}, we find the same correlations generally hold with respect to magnesium depletion. In what follows, we discuss the nature of these correlations and possible interpretations. \\subsubsection{Correlations with Hydrogen} \\label{sss:hydrogen_correlations} First, we note that magnesium depletion is clearly correlated with total hydrogen volume density $\\nHavg$. Five lines of sight (HD 27778, HD 37021, HD 37061, HD 37903, and HD 147888) have substantially larger depletions than any of the other lines of sight in this sample; HD 27778, HD 37021, HD 37061, and HD 147888 are the four densest lines of sight (in terms of $\\nHavg$), while the density of HD 37903 is in the top 20\\% of our sample. This correlation is plotted in Figure \\ref{fig:logMgIIHlognh}. Since these dense lines of sight are also some of the lines of sight that are found in both our sample and that of CLMS, we are not probing significantly higher average density and cannot expand on their conclusions regarding $\\nHavg$. We also find that magnesium depletion is correlated with total hydrogen column density. However, we conclude that this is primarily a secondary correlation due to the correlation with $\\nHavg$. However, we can attempt to analyze whether or not the correlation is independently significant. To do this, we follow the methods described in \\citet{Jenkins1986}. First, we calculate Pearson correlation coefficients between depletion and the two variables of interest (in this case $\\logHtot$ and $\\lognHavg$), as well as those two variables with each other. The partial correlation coefficient is then given by $\\rho_{12.3} = (\\rho_{12}-\\rho_{13}\\rho_{23})/[(1-\\rho_{13}^2)(1-\\rho_{23}^2)]^{-1/2}$, where the subscripts on the correlation coefficients indicate the two variables being correlated, and $\\rho_{12.3}$ is the correlation coefficient between the first two variables if the third variable is held fixed. However, what is really being calculated is $r$, the sample correlation coefficient(s), as opposed to $\\rho$, the population correlation coefficient(s). Once $r_{12.3}$ has been calculated, we examine the significance level for a $t$-test of the appropriate number of degrees of freedom (in this case, the number of data points minus three) to determine the probability that $\\rho$ is non-zero (i.e.~a true correlation exists). Using these methods, we find a 49\\% chance of the null hypothesis (i.e.~the two-sided probability that $\\rho=0$) for a correlation between Mg II depletion and $\\NHtot$ when $\\nHavg$ is held fixed. Because this technique of trivariate analysis uses Pearson correlation coefficients, a few caveats apply, namely that linear relationships and normal distributions are implicitly assumed. We used $\\logMgIIH$, $\\logHtot$, and $\\lognHavg$ in the analysis just discussed because the correlation between all combinations of those variables are stronger than when in linear form. CLMS also explored possible correlations between magnesium depletion and $\\fHmol$, but only briefly commented on the results. They concluded that any correlation was less significant than the correlation between magnesium depletion and $\\nHavg$, in that it did not as effectively discriminate between the distinct depletion levels expected in the \\citet{Spitzer1985} model. The upper right panel of Figure 9 in CLMS shows a very clear correlation between magnesium depletion and $\\fHmol$ for $\\fHmol \\gtrsim 0.1$, superimposed with a scatter plot in depletion for a smaller subset of lines with $0.01 \\lesssim \\fHmol \\lesssim 0.1$. Two additional lines of sight with $\\fHmol \\lesssim 10^{-4}$ conform to the main correlation in that they show minimal depletion. We also see a clear correlation between magnesium depletion and the molecular fraction of hydrogen, $\\fHmol$, plotted in Figure \\ref{fig:logMgIIHHf}, with depletion increasing with increasing $\\fHmol$. There is the exception of one discrepant point, HD 147888, that exhibits substantial depletion at $\\fHmol \\sim 0.1$. We note that CLMS found a larger column density for this line of sight than we do ($0.2\\dex$); however, even if we adopt the CLMS Mg II column density for this line of sight, its abundance is still $0.26\\dex$ smaller than any line of sight in our sample with $\\fHmol < 0.4$. We concur with CLMS that the correlation between magnesium depletion and $\\fHmol$ is not as rigorous as the correlation with $\\nHavg$. We examine the data using various combinations of the variables in their logarithmic (where correlations are the strongest) and linear forms. In all cases, we find that the probability of the null hypothesis (i.e.~$\\rho=0$, as discussed above) between Mg II depletion and $\\fHmol$ with $\\nHavg$ held constant is less than 5\\%; in most cases, it is $\\lesssim1$\\%. Conversely, the probability that there is no correlation between Mg II depletion and $\\nHavg$ with $\\fHmol$ held constant is less than 0.1\\%. Therefore, we conclude that while $\\nHavg$ is clearly the dominant correlation, the correlation with $\\fHmol$ has some independent significance. A question of interest is where scatter seems to be introduced and why. Between Figure 9 of CLMS and Figure \\ref{fig:logMgIIHHf} of this paper, scatter is only observed at $\\fHmol \\lesssim 0.1$. In our recent related work on Fe II \\citep{JensenFeII}, we note that Fe II depletion is also correlated with $\\fHmol$, but there are a few exceptions to the trend at both high and low values of $\\fHmol$. The most severe exceptions noted in that paper were HD 147888 and HD 164740 (with large depletions but $\\fHmol \\lesssim 0.1$) and HD 210121 (with less depletion despite $\\fHmol \\sim 0.7$). For Mg II, however, we do not see any outlying points at values of $\\fHmol$ larger than $\\sim0.1$ within either this sample or CLMS. Points such as HD 147888, however, still require explanation. \\citet{Snow1983} put forth the possibility that in some dense environments an increased average grain size, which decreases the grain surface area per unit volume, may suppress $\\Hmol$ formation (as $\\Hmol$ is thought to form on grain surfaces). This scenario was specifically discussed in the context of the $\\rho$ Oph cloud, for which there is indepedent evidence (in part, a value of $\\rv$ greater than the interstellar average of 3.1) that grain coagulation has occurred. Our main outlying line of sight, HD 147888 (with $\\rv$ of 4.06), passes through the $\\rho$ Oph cloud, so observing the combination of a dense, depleted environment with small $\\fHmol$ is not surprising in this case. Two of the other lines of sight with large depletions are HD 37021 and HD 37061. Reliable far-ultraviolet data sets do not exist for these lines of sight; therefore, they do not have measurements of the molecular hydrogen column densities or subsequently derived molecular fractions of hydrogen. However, as stated above in \\S \\ref{ss:hydrogen}, \\citet{Cartledge2001} has argued that these lines of sight have small values of $\\fHmol$ based on a lack of Cl I. These two lines of sight also have large values of $\\rv$ (5.54 and 4.23, respectively), implying a larger average grain size. The possible effect of grain size on depletions and $\\Hmol$ formation, and other interpretations, will be discussed further in \\S \\ref{sss:extinction_correlations}. Barring such outlying points, the trend of increased Mg II depletion with increased $\\fHmol$ has a relatively clear interpretation. $\\Hmol$ is formed in the same dense, dusty environments that foster large depletions. Within the context of this sample and CLMS, this seems to hold for lines of sight with $\\fHmol \\gtrsim 0.1$. While the ubiquity of $\\Hmol$ even in diffuse regions complicates the issue \\citep[see conclusions of][]{Rachford2002}, there is still a physical argument that $\\fHmol$ should be a good diagnostic of the local conditions of interstellar clouds, and therefore depletions. It is worth noting that the similar trend between iron depletion and $\\fHmol$ exhibits scatter up to $\\fHmol \\sim 0.3$ in the work of \\citet{SavageBohlin}, in addition to the outlying points mentioned above from \\citet{JensenFeII}. Whether the range in $\\fHmol$ over which there is scatter in the abundances is truly different for magnesium and iron or is simply a selection effect is unclear. \\subsubsection{Correlations with Extinction and Reddening Parameters} \\label{sss:extinction_correlations} We find that the depletion of Mg II is correlated to both selective extinction, $\\ebv$, and total visual extinction, $\\av$. However, there is significant scatter in these correlations. Because these are integrated line of sight parameters, it makes sense to divide by line-of-sight pathlength. Both $\\ebv$ and $\\av$ are strongly correlated with $\\NHtot$, and both are thought to be rough measures of the total dust column density. Therefore, $\\ebvdist$ and $\\avdist$ should be strongly correlated with $\\nHavg$ and be approximations of the total dust volume density. When we look for correlations between magnesium depletion and $\\ebvdist$ and $\\avdist$ we find that the correlations are substantially increased when compared to the integrated line of sight parameters. Therefore, we can say, with reasonable confidence, that magnesium depletion is increased in increasingly dusty environments. The correlations with $\\ebvdist$ and $\\avdist$ are plotted in Figures \\ref{fig:logMgIIHebv_dist} and \\ref{fig:logMgIIHav_dist}. As with the correlation between depletion and $\\fHmol$, we examine partial correlation coefficients to determine the independent significance of these correlations. The partial correlation coefficient between Mg II depletion and $\\log{\\ebvdist}$ with $\\lognHavg$ held fixed implies that the probability of the null hypothesis is less than 6\\%. The same partial correlation coefficient with $\\log{\\avdist}$ implies that there is less than a 1\\% chance of the null hypothesis. (Note that both probabilities are two-sided to a $t$-distribution.) We consider these variables in their logarithmic forms because these are the versions of the variables that exhibit the strong correlations (for all combinations of the variables in question). However, if the variables are considered in non-logarithmic forms, the probability of the null hypothesis generally increases. Again we note that this type of trivariate statistical measure implicitly assumes that the correlation is linear with normally distributed scatter. As with the correlation between depletion and $\\fHmol$, we conclude that while these correlations do not improve on $\\nHavg$ as a predictor of depletions, this is limited evidence that they are significant in their own right. We have briefly explored the possibility that the additional (or ``missing'') magnesium that is depleted from these lines of sight is found in the form of gas-phase Mg I in regions that are presumably shielded from radiation by dust. While the STIS data do not cover Mg I absorption lines in many cases, our results indicate that gas-phase Mg I column densities are far too small to account for the order-of-magnitude increase in depletion seen in gas-phase Mg II. Therefore, it likely that the missing gas-phase magnesium is tied up in the additional grains found in these environments, supporting the conclusion above that the correlations between Mg II depletion and the parameters $\\log{\\ebvdist}$ and $\\log{\\avdist}$ are physically significant. It is also worth noting that in no case do we see Mg/H less than $\\sim1\\ppm$, even in the densest environments, with values of $\\ebvdist$ and $\\avdist$ several times larger than the average of the sample. This suggests that these lines of sight are not probing what might be considered ``translucent clouds'' \\citep[though some may be ``translucent lines of sight''; see][]{SnowMcCall}. The Mg II abundance is plotted against the ratio of total visual to selective extinction, $\\rv \\equiv \\av / \\ebv$, in Figure \\ref{fig:logMgIIHRv}. The plot shows significant scatter; however, some statistical measures show the possibility of a slight correlation. A Pearson correlation coefficient between $\\logMgIIH$ and $\\rv$ implies that the probability of the null hypothesis is about 31\\%. A Spearman's $\\rho$ rank correlation coefficient, which does not depend on the functional form assumed (including whether variables are considered linearly or logarithmically) beyond assuming that the correlation is either monotonically increasing or monotonically decreasing, shows a negative correlation (decreasing abundance/increasing depletion as $\\rv$ increases) and is significant to approximately 1.3-$\\sigma$. Therefore, there is evidence of a possible slight correlation between depletion and $\\rv$. However, this is far from a certain conclusion. Significant selection effects are also a possibility, as the correlations are dominated by some of the points that have greater depletion and large values of $\\rv$. In fact, if these lines of sight are excluded, the trend begins to reverse toward a positive correlation between increasing Mg II abundance and increasing $\\rv$. In general, we conclude that $\\rv$ is a poor predictor of depletions; as one anecdotal counterexample, HD 91597 has a very large value of $\\rv=4.9$ but does not exhibit particularly large Mg II depletion. However, there are a few lines of sight that present interesting interpretive challenges where the value of $\\rv$ may provide insight. As discussed above, the possibility of large magnesium depletion but small $\\fHmol$ exists for three lines of sight: HD 37021, HD 37061, and HD 147888. In the latter case the effect is clear, while in the former two cases the small value of $\\fHmol$ is merely inferred. What is interesting, as noted above, is that these three lines of sight all have values of $\\rv>4$ which is a fairly significant deviation from the interstellar average of 3.1. Because large grains contribute to $\\av$ (i.e.~grey extinction) but less so to $\\ebv$, $\\rv$ is thought to be correlated to average grain size. Explaining why depletion should increase in a line of sight with large grains is difficult. As mentioned above in our discussion of $\\Hmol$ and iron depletion, increased grain size decreases dust surface area per unit volume, and therefore reduces $\\Hmol$ formation rates. However, decreased surface area per unit volume also implies a reduction in rates of sticking between dust grains and gas-phase atoms. It seems we can reasonably conclude that the large values of $\\rv$, i.e.~the larger average grain populations, are not responsible for the large depletions by way of atoms and ions sticking to the grains. One possibility is that the large depletions are instead ``locked in'' prior to grain coagulation. Another possibility is the effect of a high-radiation field: this is known for the line of sight toward HD 147888 ($\\rho$ Oph D) as well as HD 37021 and HD 37061 which are in Orion (radiation is presumed to be responsible for the relative lack of Cl I, and thus also $\\Hmol$, as mentioned in \\S \\ref{ss:hydrogen}). However, \\citet{Snow1983} argues that the increased radiation is unlikely to be entirely responsible for the low $\\fHmol$ in the $\\rho$ Oph cloud. Whether or not this is the case for HD 37021 and HD 37061 is unclear. More details about the radiation field and the exact nature of the grain population (we have only considered the crude measure of $\\rv$) are probably necessary to fully understand these lines of sight. \\subsubsection{Anticorrelation with Distance} \\label{sss:distance_anticorrelation} We find that magnesium depletion is generally anticorrelated with distance to the background star, that is, line of sight pathlength; this relationship is shown in Figure \\ref{fig:logMgIIHdist}. Depletion decreases by nearly an order of magnitude between very short lines of sight and those up to about 2 kpc or so, and then is relatively constant (to within about 0.3-0.4$\\dex$) out to about 6 kpc. As we concluded for a similar anticorrelation seen between iron depletion and distance \\citep{JensenFeII}, the long pathlengths are likely sampling a variety of cloud conditions, resulting in the constant depletion for long-pathlength lines of sight. On the other hand, given the comparable hydrogen column densities of all the lines of sight in this study ($\\logHtot \\approx21-22$), the shorter lines of sight are generally the denser lines of sight. \\subsubsection{Spatial Variations} \\label{sss:spatial_variations} We find one very interesting correlation with Galactic location: the five stars with the largest depletions reside at higher Galactic latitudes of $|b|>15^{\\circ}$. However, with pathlengths of less than 1 kpc, these lines of sight are still primarily in the Galactic disk. When we analyze magnesium depletions against the height from the center of the Galactic disk, $z=r \\sin{b}$, we do not see a strong correlation. The variation with respect to Galactic latitude is most likely a coincidence, given that these are some of the densest and most reddened lines of sight. We do not see any other evidence of significant spatial variations. \\subsection{Mg/H of Galactic Stars} \\label{ss:stellarMgH} In the last several years, three major papers have attempted to analyze the cosmic abundance ``standards'' in the ISM through studies of stellar abundances and meteoritic abundances---\\citet{SnowWitt}, \\citet{SofiaMeyer}, and \\citet{Lodders}. The importance of these standards is to compare them with the observed gas-phase abundances and infer an absolute value for depletions---and therefore absolute values for the amount of these elements in phases other than atomic gas, i.e.~dust grains and molecules. Of the major elements relevant to dust, the element with the best determined cosmic abundance is iron. The four major measurements of the cosmic Fe/H ratio---solar, B stars, F and G stars, and CI chondrites---all agree very closely, largely within the errors. The situation is somewhat more complex for other elements. Carbon and oxygen show apparent overabundances in the Sun compared to F and G stars (whether or not the solar and F/G star abundances potentially agree within the errors depends on the choice of solar abundances, regarding which there is still some uncertainty), while B stars show relative deficits in these abundances compared to the Sun and other F and G stars. The chondritic abundances of C and O are even smaller. Nitrogen seems to be somewhat less abundant in B stars than in the Sun, though the two are reconciliable within the errors; F and G nitrogen abundances are generally unknown, and the chondritic abundances are substantially lower. Silicon seems to be most abundant in F and G stars, slightly less abundant in the Sun, and about half as abundant in B stars. The errors, however, do not rule out agreement between all three measurements. However, the chondritic abundance tightly matches the solar abundance. Both \\citet{SofiaMeyer} and \\citet{Lodders}, cite \\citet{Holweger} for the solar abundance of magnesium, $\\logMgH=-4.46$; \\citet{SnowWitt} report a slightly older value from \\citet{AndersGrevesse} of $\\logMgH=-4.42$, though these values are consistent within the errors. The chondritic abundances in \\citet{Lodders} of $\\logMgH=-4.44$ are also very consistent with these solar values. The discrepancy arises when various stellar abundances are considered. Both \\citet{SnowWitt} and \\citet{SofiaMeyer} found significantly smaller abundances of Mg for B stars. \\citet{SnowWitt} found $\\logMgH=-4.63$ for field B stars and $\\logMgH=-4.68$ for cluster B stars; \\citet{SofiaMeyer}, making no distinction between cluster and field stars, found $\\logMgH=-4.64$ for all B stars. Though there is marginal agreement within the very large errors in these numbers, the B star abundances are $\\approx60\\%$ smaller than the solar abundances. \\citet{SnowWitt} and \\citet{SofiaMeyer} also disagree on the Mg abundance in F and G stars ($\\logMgH=-4.52$ and $-4.37$, respectively) due most likely to \\citet{SofiaMeyer} restricting their sample to stars with ages of $\\leq2$ Gyr. Again, however, these values have relatively large errors and are reconciliable with the B star abundances, though just barely. What effect does the choice of a cosmic magnesium abundance have for the implied dust-phase abundances to be used in dust models? The differences between the cosmic abundances just discussed leads to nontrivial differences in the dust-phase abundances. Our weighted interstellar average of Mg II/H is $2.7\\pm0.1\\ppm$ (parts per million), though the median value in our 44 lines of sight is somewhat larger at $6.2\\ppm$. Taking the extremes of the above numbers, anywhere between $\\sim20$ and $\\sim40\\ppm$ of Mg is available for creating dust. Examining the various models compared in Table 3 of \\citet{SnowWitt}, we find that most models require much more Mg than the lower value of $\\sim20$ implied by a B star abundance standard. That a B star abundance is less likely to represent the cosmic abundance was also found by \\citet{ZDA2004}, who had more difficulty fitting dust models to observations using the dust-phase abundances implied by assuming B star abundances as the cosmic standard. In fact, this is true even though \\citeauthor{ZDA2004} assumed $\\approx0$ ppm of magnesium to be in the gas-phase. If the few ppm of magnesium in the gas-phase as measured by this paper and CLMS were included, the \\citet{ZDA2004} fits would become even more strained (the best fits for B star abundances were at the limit of the error in those abundances and inferior to the fits obtained using other abundances). Therefore, the major conclusion that we can make regarding cosmic abundances and the incorporation of magnesium into dust is to add to the evidence that B star abundances, despite B stars being younger and therefore potentially good tracers of the current ISM, are a poor cosmic standard. Whether or not the solar or an F and G star abundance standard for Mg is a better fit is a test that is too sensitive for us to comment on, given the uncertainties in those abundances. We have analyzed the abundance of Mg II in 44 lines of sight. Our study does not probe substantially larger average hydrogen volume densities than previously observed by CLMS; therefore, we observe the same correlations between Mg II and the $\\nHavg$ and $\\ebvdist$. We also note a correlations between magnesium depletion and $\\avdist$, a different measure of dust density. Correlations between $\\NHtot$ and the reddening and extinction parameters $\\ebv$ and $\\av$ mean that correlations between Mg II depletion and dust density measures are expected. However, these latter correlations, while not strong than the correlation between depletion and $\\nHavg$, show some evidence of being significant even at fixed $\\nHavg$ and should be more directly related to the line-of-sight dust content. We also note a correlation between magnesium depletion and $\\fHmol$ in our data; combined with the results of CLMS, this correlation seems to be valid for $\\fHmol \\gtrsim 0.1$ but not at smaller $\\fHmol$. A question that is related to the trend with $\\fHmol$ is why so little $\\Hmol$ forms in certain high-density lines of sight. Our results are consistent with the \\citet{Snow1983} suggestion that the reduced grain surface area per unit volume of large grains plays a role in reducing $\\Hmol$ formation rates. For similar reasons, we can conclude that the grain coagulation probably occurs after depletions are already ``locked into'' the dust, rather than depletion of gas-phase atoms onto grain surfaces." }, "0710/0710.1913_arXiv.txt": { "abstract": "We report on the H$_2$O maser distributions around IRAS 22480+6002 (=IRC+60370) observed with the Japanese VLBI Network (JVN) at three epochs spanning 2 months. This object was identified as a K-type supergiant in 1970s, which was unusual as a stellar maser source. The spectrum of H$_2$O masers consists of 5 peaks separated roughly equally by a few km s$^{-1}$ each. The H$_2$O masers were spatially resolved into more than 15 features, which spread about 50 mas along the east--west direction. However, no correlation was found between the proper motion vectors and their spatial distributions; the velocity field of the envelope seems random. A statistical parallax method applied to the observed proper-motion data set gives a distance of $1.0\\pm 0.4$ kpc for this object, that is considerably smaller than previously thought. The distance indicates that this is an evolved star with $L\\sim 5800\\ L_{\\odot}$. This star shows radio, infrared, and optical characteristics quite similar to those of the population II post-AGB stars such as RV Tau variables. ", "introduction": "\\label{sec:introduction} H$_2$O maser emission has been observed in circumstellar envelopes of evolved stars such as O-rich Mira variables and OH/IR stars with large mass loss rates of $\\dot{M} \\geq 10^{-7}M_{\\odot}$~yr$^{-1}$ \\citep{rei81,eli92}. Most of these stars are asymptotic giant branch (AGB) stars or red supergiants both with the spectral type M, with a few exceptions for transient stars at pre-planetary nebula phase (or supposedly a few pre-main sequence stars such as Ori KL; \\cite{mor98}). For the central stars with spectral types earlier than M, UV radiation from stellar chromosphere eventually dissociates most of molecules (except CO) in the inner envelope (e.g., \\cite{wir98}). Therefore, H$_2$O (or SiO) masers are usually not expected for these stars, except for the case that the molecules in dense circumstellar clumps shield themselves from UV radiation. In fact, H$_2$O masers found in a young planetary nebula \\citep{mir01,sue07} must be such an exceptional case. OH masers have been found in yellow hypergiants with spectral types F and G (such as IRC+10420 and V1427 Aql) \\citep{gig76,ned92,hum02}. However, H$_2$O and SiO masers have never been detected in these objects \\citep{nak03}, though thermal emission of a few other molecules have been observed in the outer circumstellar shell \\citep{cas01,tey06}. \\begin{table*}[ht] \\caption{Status of the telescopes, data reduction, and resulting performances in the individual epochs of the JVN observations.}\\label{tab:status} \\begin{center} \\footnotesize \\begin{tabular}{lccccccc} \\hline \\hline & Epoch in & & & Reference & 1-$\\sigma$ level & Synthesized & Number of \\\\ Observation & the year & Duration & Used & velocity\\footnotemark[2] & noise & beam\\footnotemark[3] & detected \\\\ code & 2005 & (hr) & telescopes\\footnotemark[1] & (km s$^{-1}$) & (Jy beam$^{-1}$) & (mas) & features \\\\ \\hline r05084b \\dotfill & March 25 & 7.3 & MZ, IR, OG, IS, KS, NB\\footnotemark[4] & $-52.3$ & 0.22 & 1.7$\\times$1.6, $-$37$^{\\circ}$ & 20 \\\\ r05116b \\dotfill & April 26 & 7.3 & MZ, IR, OG, IS\\footnotemark[5], KS, NB & $-52.0$ & 0.15 & 3.8$\\times$2.0, $-$14$^{\\circ}$ & 17 \\\\ r05151a \\dotfill & May 31 & 8.1 & MZ, OG\\footnotemark[5], IS\\footnotemark[5], KS, NB & $-$52.6 & 0.15 & 3.2$\\times$2.8, $-$66$^{\\circ}$ & 14 \\\\ \\hline \\end{tabular} \\end{center} \\footnotemark[1] Telescopes that were effectively operated and whose recorded data were valid: MZ: the VERA 20~m telescope at Mizusawa, IR: the VERA 20~m telescope at Iriki, OG: the VERA 20~m telescope at Ogasawara Is., IS: the VERA 20~m telescope at Ishigakijima Is., KS: the NiCT 34-m telescope at Kashima, NB: the NRO 45-m telescope at Nobeyama. \\\\ \\footnotemark[2] Velocity channel used for the phase reference in data reduction. \\\\ \\footnotemark[3] The synthesized beam made in natural weight; major and minor axis lengths and position angle. \\\\ \\footnotemark[4] Ceasing operation for 2.5~hr due to strong winds and pointing correction. \\\\ \\footnotemark[5] High system temperature ($>$300~K) due to bad weather conditions. \\end{table*} \\ \\\\ The optical counterpart of IRAS 22480+6002 ($=$AFGL~2968, or IRC$+$60370) was identified as a K-type supergiant (K0Ia; \\cite{hum74}, or K4.5Ia; \\cite{faw77}). Therefore, the detections of H$_2$O and SiO masers \\citep{han98,nym98} were surprising. Though a search for OH 1612 MHz emission was negative \\citep{les92}, CO emission was detected toward this star \\citep{jos98}. From the CO $J=2$--1 line profile, the systemic stellar velocity and the expansion velocity of this star were obtained to be $V_{\\rm lsr}$=$-49.3$ km s$^{-1}$ and $V_{exp}=26.4$ km~s$^{-1}$, respectively \\citep{jos98, gro99}. They are consistent with those obtained from the H$_2$O\\ and SiO maser spectra, and the expansion velocity of the envelope of this star is typical for OH/IR stars. The radial velocity gives a kinematic distance of 5.0~kpc. It suggests a large luminosity $L_{\\ast}=$140 000~$L_{\\odot}$ of the central star \\citep{gro99}, but it is consistent with the supergiant interpretation of this object. A blue nearby star, a B5II star, is seen by about 12$''$ east of this object. Though it is cataloged as a visual binary \\citep{wor97}, a physical association of this object with the maser source is questionable because of the large velocity difference of about 40 km s$^{-1}$ \\citep{hum74}. \\citet{win94} gave a new spectral classification of M0I for IRAS 22480+6002 from the low-resolution spectrum between 6000 and 8800 A, which was significantly different from the previous type assignment of this star. For a long-period variable, optical spectral classification may vary with light variations. However, this star has not been reported as a variable star, though it is optically not very faint ($V\\sim 8.3$). In this work, we report three-epoch VLBI observations of H$_2$O\\ masers of IRAS 22480+6002 to rectify the entangled situation associated with this object. From the spatio-kinematics of the masers, we diagnose a probable anomaly of a hot wind from the K-type star. We estimated the distance to this star using the statistical parallax method based on the proper motion data of H$_2$O masers. Our result gives a much smaller distance for this star than previously thought. The new estimation of the distance demands to reconsider various properties of this star. Based on the arguments presented in section 3, we conclude that this star is a population II post-AGB star. ", "conclusions": "\\begin{table*}[ht] \\caption{Parameters of the H$_2$O maser features identified by proper motion toward IRAS 22480$+$6002.} \\label{tab:pmotions} \\begin{center} \\begin{tabular}{lrrrrrrrrrrr} \\hline \\hline Feature\\footnotemark[1] & \\multicolumn{2}{c}{Offset} & \\multicolumn{4}{c}{Proper motion\\footnotemark[2]} & \\multicolumn{2}{c}{Radial motion\\footnotemark[3]} & \\multicolumn{ 3}{c}{Peak intensity at 3 epochs} \\\\ & \\multicolumn{2}{c}{(mas)} & \\multicolumn{4}{c}{(mas yr$^{-1}$)} & \\multicolumn{2}{c}{(km s$^{-1}$)} & \\multicolumn{ 3}{c}{(Jy beam$^{-1}$)} \\\\ & \\multicolumn{2}{c}{\\hrulefill} & \\multicolumn{4}{c}{\\hrulefill} & \\multicolumn{2}{c}{\\hrulefill} & \\multicolumn{ 3}{c}{\\hrulefill} \\\\ & $\\Delta$R.A. & $\\Delta$decl. & $\\mu_{x}$ & $\\sigma(\\mu_{x})$ & $\\mu_{y}$ & $\\sigma(\\mu_{y})$ & $V_{z}$ & $\\Delta V_{z}$\\footnotemark[4] & Epoch 1& Epoch 2& Epoch 3 \\\\ \\hline 1 \\ \\dotfill \\ &$ -6.56$&$ -10.69$&$ 4.20$& 1.46 &$ 5.86$& 1.40 &$ -59.78$& 0.26 & 0.19 & 0.17 & ... \\\\ 2 \\ \\dotfill \\ &$ 0.00$&$ 0.00$&$ 0.00$& 0.36 &$ 0.00$& 1.29 &$ -58.48$& 2.74 & 2.33 & 2.72 & 3.48 \\\\ 3 \\ \\dotfill \\ &$ -45.23$&$ 7.26$&$ -0.87$& 0.42 &$ 0.51$& 0.66 &$ -55.01$& 2.04 & 3.77 & 2.47 & 2.42 \\\\ 4 \\ \\dotfill \\ &$ -4.95$&$ 7.17$&$ 0.25$& 0.29 &$ 0.47$& 1.44 &$ -54.33$& 1.05 & 0.41 & 0.53 & 0.73 \\\\ 5 \\ \\dotfill \\ &$ -26.70$&$ -2.86$&$ -1.03$& 0.97 &$ -0.72$& 0.61 &$ -52.34$& 3.37 & 8.11 & 8.73 & 8.16 \\\\ 6 \\ \\dotfill \\ &$ -0.37$&$ 1.50$&$ 3.96$& 1.29 &$ 0.86$& 1.91 &$ -50.37$& 0.56 & 0.53 & 0.73 & 0.49 \\\\ 7 \\ \\dotfill \\ &$ 0.16$&$ 2.74$&$ 0.05$& 0.61 &$ 0.98$& 1.56 &$ -49.29$& 2.18 & 8.44 & 7.96 & 6.20 \\\\ 8 \\ \\dotfill \\ &$ -0.04$&$ 13.96$&$ 0.90$& 0.67 &$ -0.29$& 0.78 &$ -47.88$& 1.33 & 1.92 & 1.69 & 1.03 \\\\ 9 \\ \\dotfill \\ &$ -34.92$&$ -7.63$&$ -1.95$& 1.66 &$ -3.07$& 2.97 &$ -47.39$& 0.52 & 0.25 & 0.26 & ... \\\\ 10 \\ \\dotfill \\ &$ -48.23$&$ 11.08$&$ 2.27$& 1.18 &$ 0.71$& 0.88 &$ -47.24$& 0.56 & 0.24 & 0.20 & 0.18 \\\\ 11 \\ \\dotfill \\ &$ -2.24$&$ 8.18$&$ 0.06$& 1.10 &$ -0.16$& 0.64 &$ -46.17$& 0.84 & 0.41 & 0.28 & 0.26 \\\\ 12 \\ \\dotfill \\ &$ -1.02$&$ 13.29$&$ 0.06$& 0.41 &$ 0.56$& 0.74 &$ -45.37$& 1.97 & 3.34 & 3.40 & 2.44 \\\\ 13 \\ \\dotfill \\ &$ 0.13$&$ 7.77$&$ -1.86$& 3.19 &$ 0.26$& 1.32 &$ -44.40$& 0.77 & 0.24 & 0.33 & 0.32 \\\\ \\hline \\end{tabular} \\end{center} \\noindent \\footnotemark[1] H$_2$O maser features detected toward IRAS 22480+6002. The feature is designated as IRAS 22480+6002:I2007 {\\it N}, where {\\it N} is the ordinal source number given in this column (I2007 stands for sources found by Imai et~al. and listed in 2007). \\\\ \\footnotemark[2] Relative value with respect to the motion of the position-reference maser feature: IRAS 22480+6002:I2007 {\\it 2}. \\\\ \\footnotemark[3] Relative value with respect to the local stand of rest. \\\\ \\footnotemark[4] Mean full velocity width of a maser feature at half intensity. \\end{table*} \\subsection{Spatial distribution and proper motions of maser features} Figure \\ref{fig:I2248_spectrum} shows cross-power spectra of the H$_2$O masers of IRAS 22480+6002. The H$_2$O maser emission spread in a velocity range of 15 km~s$^{-1}$, which is typical for Mira-type AGB stars (e.g., \\cite{tak94}). Five spectral peaks were seen in roughly equal separations of 2--3 km~s$^{-1}$; the second highest peak was near the systemic velocity ($V_{\\rm lsr}=-49.3$ km s$^{-1}$). The correlated powers of these peaks equally increased by a factor of two during 2 months in our observing run, except for the second peak for which the intensity increased only by about 20\\%. However, the peak flux densities of individual features were found not to vary much (Table 2). This fact indicates that extended emissions were partially resolved in the shortest baseline between NRO and NICT (197.4 km), but resolved-out in the longer baselines. % Correlated flux densities were estimated to be about 30--40\\% of the total-power intensities. Figure \\ref{fig:1st-epoch} shows the distribution of maser features at the first epoch. The extent of 50 mas corresponds to 50~AU ($D$/1 kpc), which is somewhat larger than those seen around Mira variables at $D=1$ kpc (e.g., \\cite{bow94}). In this figure, one of the low-velocity (blue-shifted) components ($V_{\\rm lsr}=-58.48$ km s$^{-1}$: the position reference) is located at the origin and the other low velocity components (in grey and blue colors) are located both near the eastern and western edges. Many of the higher velocity (red-shifted) components (shown in yellow, orange, and red colors) fall at the eastern edge, but a few of them are scattered at the west side too. The overall distribution of water maser features are characterized by the elongation to the east-west direction. But, no clear correlation is found between the velocities and spatial positions. If the circumstellar envelope of the K supergiant interacts with the wind from the eastern BII star (though this is unlikely), maser spots and features could be aligned perpendicularly to the wind direction, i.e., in the north--south direction (e.g., \\cite{ima02b}). We find no such N--S alignment of the H$_2$O maser features. \\citet{mei99} noted that the MIR image of this star with the NASA 3-m telescope showed a northeast-southwest elongation, but concluded that it was likely to be an artifact caused by astigmatism. \\begin{figure}[htb] \\begin{center} \\FigureFile(8cm,8cm){fig2.eps} \\end{center} \\caption{Distribution of H$_2$O masers on 2005 March 25. The color code indicates the radial velocity of the feature, and the size of the filled circle indicates the flux density of the feature. Note that the $-58.48$ km s$^{-1}$ reference component is located at the origin (light blue), but is almost overlapped with the systemic-velocity component ($\\sim -49$ km s$^{-1}$) shown in green.} \\label{fig:1st-epoch} \\end{figure} We detected 14--20 H$_2$O maser features though all epochs (the last column of table \\ref{tab:status}). Note that the H$_2$O masers were persistent in velocity and in spatial distribution during the two months; 65--90\\% of the detected maser features survived during our observing run. Therefore, we identified the same maser features at three epochs relatively easily, and measured the proper motions of the individual features during two months. Table \\ref{tab:pmotions} gives the measured proper motions. Figure \\ref{fig:PM-I2248} shows the linear fits to the relative positions of the individual maser features. The fitted proper motions look significant with the second epoch contributing little for most features, which warrants our selecting the same maser features at the different epochs. % Circumstellar H$_2$O masers can amplify the radiation of the central star (for example, see the case of U Her; \\cite{vle02}). % In the present case, we may speculate that the $-52.34$ km s$^{-1}$ feature (No. 5 in Table 2 and Figure 4), which is one of the strongest components and located near the center of the maser distribution, is such a maser amplifying the steller radiation. However, it is hasty to draw any conclusion from this observation, since we have no information on the central-star position in this scale. \\begin{figure}[ht] \\FigureFile(8cm,14cm){fig3.eps} \\caption{Observed relative proper motions of H$_2$O maser features in IRAS 22480$+$6002 in R.A. (a) and decl. (b) directions. The number on the left indicates the feature name designated in Table 2. The vertical bar attached to each data point indicates the position uncertainty. The least-square-fitted line is also shown.} \\label{fig:PM-I2248} \\end{figure} \\begin{figure}[ht] \\FigureFile(8cm,6cm){fig4.eps} \\caption{Doppler velocities (colorfully displayed) and relative proper motion vectors (indicated by arrows) of H$_2$O masers in IRAS 22480$+$6002. The displayed proper motion vector is that subtracted by a velocity bias ($\\overline{\\mu _{x}}, \\overline{\\mu _{y}})=(0.97, 0.72)$ [mas yr$^{-1}$] from the original vector to cancel out the average motions of all the features. A number added for each feature with a proper motion shows the assigned one after the designated name form ``IRAS 22480$+$6002: I2007\". The map origin is set to the location of the feature IRAS 22480$+$6002: I2007 {\\it 2}.} \\label{fig:I2248-velocity} \\end{figure} Figure \\ref{fig:I2248-velocity} shows the proper motion vectors of the individual H$_2$O maser features. Note that the largest two proper-motion vectors at the lower left and lower right, i.e., features 1 and 9 of the $V_{\\rm lsr}=-59.78$ and $-47.39$ km s$^{-1}$ components, respectively, were determined by two-epoch detections, so that they are slightly inaccurate. The proper motions of all other features with 3-epoch detections are within a few mas per year (relative to the reference component at $V_{\\rm lsr}=-58.48$ km s$^{-1}$). We cannot find any systematic trend of motions in this diagram. For example, features 3 and 10 at the western edge move in opposite directions, and features 6 and 13 at the eastern edge also move in opposite directions. Figure \\ref{fig:expansion} shows the RA-offsets and $\\mu_x$ plots against $V_{\\rm lsr}$. The ellipse is a plot of expected offset and proper motion from a thin spherical-shell model with a constant velocity (in the Right Ascension direction because the maser features are spread mainly in this direction). If the shell model is correct, all of the maser features should fall between these ellipses. However, the right panel does not show such a tendency. Rather, the observed points seem to distribute randomly. The randomness of the proper motions may partially originate from the large random errors in the position measurements. % In order to check this issue, we made a Monte Carlo simulation of the 3-epoch proper motion fitting with the same positional uncertainties but without real motions (i.e., position jitters only due to the measurement errors). We obtained the mean velocity dispersions ($0.79\\pm 0.25$ mas yr$^{-1}$, $0.80\\pm 0.20$ mas yr$^{-1}$) for the 13 proper motions in R.A. and Dec. directions for the present case from the simulations; here, the number after the '$\\pm$' sign is a standard deviation of dispersions obtained in the simulations. The observed velocity dispersions (1.95 mas yr$^{-1}$, 1.93 mas yr$^{-1}$), are significantly larger than the simulated mean dispersions (more than $4 \\sigma$). Therefore, the observed proper motions are substantially real motions of masing features. \\begin{figure}[ht] \\FigureFile(8cm,6cm){fig5.eps} \\caption{Plot of relative R.A. offset (left) and proper motion (right) against radial velocity. The filled and unfilled circles indicate the three-epoch and two-epoch detections. The large and small ellipses in both panels indicate the position and proper motion curves expected from thin spherical shell models (one-dimensional in the R.A. direction) with a constant expansion velocity of 12.5 km s$^{-1}$ and a radius of $3.5\\times 10^{14}$ cm, and with 7 km s$^{-1}$ and $8\\times 10^{13}$ cm, respectively both at a distance of 0.9 kpc. The observed points in the right panel do not fit to these ellipses. } \\label{fig:expansion} \\end{figure} The observed random motions may originate from the intrinsic random ballistic motions of matters ejected from or infalling into the atmosphere of the central supergiant. This may be a characteristic of mass outflow of a supergiant originating from the extended atmosphere which is considerably turbulent \\citep{lev05,jos07}. The motion of 1 mas yr$^{-1}$ corresponds a transverse motion of $\\sim 4.7 (D/$kpc) km s$^{-1}$. In order to obtain the distance, we applied the statistical parallax method to the obtained maser proper motions; % for example, see \\citet{sch81}. % Assuming random motions of maser features, we obtained a velocity dispersion in radial motion (with respect to the average velocity of maser features, $V_{\\rm lsr}=-50.6$ km s$^{-1}$) to be $\\sigma_v \\simeq 5.0$ km s$^{-1}$, which is smaller than the outflow velocity estimated from CO emission. The dispersion in the maser proper motions can be obtained by subtracting the dispersion involved in the measurements; % see equation (3) of \\citet{sch81}. % We obtain $\\sigma _{\\mu} \\sim 1.40$ mas yr$^{-1}$ and get a distance to IRAS 22480+6002 to be $D=\\sigma_{v}/\\sigma_{\\mu}= 0.76 \\ (\\pm 0.25)$ kpc. The formal uncertainty involved in the distance estimation was computed using equation (4) of \\citet{sch81}. If we exclude the largest two proper-motion features with two-epoch detections from the sample, we get the distance $1.02 \\ (\\pm 0.38) $ kpc. Later on, we adopt this distance for IRAS 22480+6000, because large motions detected by two-epoch observations are somewhat dubious. Note that this distance is derived based on the assumption that the velocity field of masers is random and isotropic, and that the proper motions appeared in maser features are real motions of gas clumps. The distance 1.0 kpc gives a radius of water maser shell approximately $3.7\\times 10^{14}$ cm, which is compatible with the radii of water maser shells of miras, but considerably smaller than those of M-supergiants \\citep{yat94,cot04}. Though the obtained distance still involve a considerable uncertainty, it excludes the possibility of a very large distance of 5 kpc (a kinematic distance). The luminosity of this star is re-evaluated to be $5.8 \\times 10^3~L_{\\odot}$ (reestimated from \\cite{gro99}). It is considerably small for a supergiant. However, from the distance of 1.0 kpc, we can compute the absolute V magnitude of this star from 2MASS K magnitude ($K=2.8$) using $V-K=3.7$ (for K5III; \\cite{zom90}), and with reddening corrections, we get $M_V\\sim -4.4$. This value falls near the absolute magnitude of K5Ib (or M0Ib) \\citep{zom90}. Therefore, the luminosity is still in a range of supergiants. The radial velocity of $\\sim -50$ km s$^{-1}$ is typical for young objects in the Perseus spiral arm in the direction of this star (for example, see \\cite{sit03}). If we take into account the large uncertainties involved in the obtained distance, we cannot completely exclude the possibility that IRAS 22480+6002 belongs to the Perseus arm at $D\\sim 3.0$ kpc at $l=108^{\\circ}$ \\citep{xux06}. However, there are several other bright stellar maser sources with similar radial velocities in the same direction, e.g., CU Cep ($-50$ km s$^{-1}$), IRC+60374 ($-52$ km s$^{-1}$), and AFGL 2999 ($-50$ km s$^{-1}$). Luminosity distances to these stars are inferred to be smaller than 3 kpc from their high IRAS flux densities. In addition, MY Cep is an M supergiant with $V_{\\rm lsr}=-56$ km s$^{-1}$ in the star cluster NGC 7419. The distance to this cluster has been well estimated to be about 2.3 kpc from luminosities of member stars of the cluster (e.g., see \\cite{bea94,sub06}). These example indicates that the stars with $V_{\\rm lsr}\\sim -50$ km s$^{-1}$ do not necessarily belong to the Perseus arm, but they may be located much closely. Because the radial velocity expected by the galactic rotation is only $\\sim -10$ km s$^{-1}$ at 1 kpc at $l=108^{\\circ}$, and because the radial-velocity dispersion of stellar maser sources is as small as $\\sim 25$ km s$^{-1}$ at the solar neighborhood (see Appendix 2 of \\cite{deg05}), IRAS 22480+6002 is possibly kinematically anomalous. \\begin{table*}[ht] \\begin{center} \\caption{Comparison of the catalogued positions of IRAS 22480$+$6002.} \\label{tab:positions} \\footnotesize \\begin{tabular}{llllllll} \\hline \\hline Catalog & Band & Assignment & epoch & R.A.(J2000) & decl..(J2000) & error & flux density \\\\ & & & & \\ h \\ m \\ s & \\ \\ ${\\circ} \\ \\ ' \\ \\ ''$ & $''$ & or magnitude \\\\ \\hline IRAS & MIR & 22480+6002 & 1983 & 22 49 59.2 & +60 17 55 & $11''\\times 5''$ (19$^{\\circ}$) & $F_{\\rm 12}=142$ Jy \\\\ MSX6 & MIR & G108.4255+00.8939 & 1995 & 22 49 58.89 & +60 17 56.8 & 0.3 & $F_{\\rm C}=123$ Jy\\\\ & & & & & & & \\\\ 2MASS & NIR & 2495897+6017567 & 1997 & 22 49 58.97 & +60 17 56.8 & 0.29 & K=2.78 \\\\ GSC1.2 & optical& 0426500695 & 1954 & 22 49 59.43 & +60 17 55.8 & 0.3 & R=12.29 \\\\ GSC2.2 & optical& N012302336407 & 1989.6 & 22 49 58.900 & +60 17 57.17 & 0.3 & B=12.29 \\\\ & & & & & & & \\\\ USNO-B1.0 & optical& 1502-0356025 & 1971.7 & 22 49 59.75 & +60 17 56.7 & (0.7, 1.0) & R=8.87 \\\\ USNO-B1.0 & optical& 1502-0356023 & 1979.7 & 22 49 59.44 & +60 17 55.9 & (0.7, 0.2) & R=8.73 \\\\ USNO-B1.0 & optical& 1502-0356019 & 1979.7 & 22 49 59.15 & +60 17 57.5 & (0.5, 0.7) & B=12.63 \\\\ & & & & & & & \\\\ JVN (this work) & radio & 22480+6002 & 2005.5 & 22 49 58.876 & +60 17 56.65 & $0.1''$ & $F_{\\rm H_2O}\\sim 8$ Jy \\\\ \\hline \\hline \\end{tabular} \\end{center} \\end{table*} \\subsection{Past optical/infrared data of IRAS 22480+6002 (=J22495897+6017568).} Though this star is relatively bright at optical wavelengths ($V\\sim 8.30$; Tyco Input catalog), the star was not recorded in major optical catalogs, for example, not in Henry Draper (HD) Catalogue, The Hipparcos and Tycho Catalogue, nor the General Catalog of Variable Stars, possibly because of confusion by the nearby B5II star (TYC 4265-870-1; $V\\sim 10.74$), located by about 12$''$ east. This was involved in The Washington Double Star Catalog\\footnote{available at http://ad.usno.navy.mil/wds/wdstext.html.}, giving a separation of 10.9$''$ in 1901 and 12.0$''$ in 2006 with a small position angle variation (by $\\sim 7^{\\circ}$) to the B star. From this data, we obtain the proper motion of 17 mas yr$^{-1}$ to the west for IRAS 22480+6002 relative to this B star. As noted by \\citet{hum74}, this B5II star is probably not a binary counterpart because of the large radial velocity difference. The ACT Reference Catalog gave a very small proper motion of this B5II star (less than 3 mas yr$^{-1}$); though The Hipparcos and Tycho Catalogue gave a large proper motion in declination due to position uncertainty, but this was corrected in ACT catalog. We also checked the past catalogs recording the position of this star and summarized the results in table 3.\\footnote{ all the data except JVN are available in the VizieR database ({\\it http://vizier.nao.ac.jp/viz-bin/VizieR}).} The GSC 1.2 catalog (which remeasured the POSS1 plate taken in 1950s) gave a different position by about 4.2$''$, which leads a large proper motion of 84 mas yr$^{-1}$ if compared with GSC 2.0. This value is much larger than the above-mentioned proper motion computed from the Washington Double Star Catalog, though the proper motion vectors are roughly in the same direction. We believe the direct measurements of binary separation gives better values. Therefore, we adopt the proper motion of 17 mas yr$^{-1}$ for this star, and get $U_0=-71$ km s$^{-1}$ and $V_0=-29$ km s$^{-1}$ for IRAS 22480+6002. This motion is considerably peculiar for a population I disk star. It is likely that IRAS 22480+6002 belongs to one of kinematical streaming groups of stars as population II G and K giants \\citep{fam05}. In the past, OH and H$_2$O masers have been found in a few planetary and preplanetary nebulae \\citep{zil89,mir01,sue07}, where central stars of these objects have spectral types earlier than M. These masers are a remnant of circumstellar material which was ejected at the AGB phase of the central star. The molecules responsible for masers are eventually to be dissociated. In contrast, \\citet{fix84} found OH 1665/1667 MHz emission toward several warm stars as RV Tau variables with spectral type of G and K, but so far only one case (TW Aql, a semi-regular variable of K7III) was confirmed to be a circumstellar maser \\citep{pla91}. SiO masers, which are emitted within a few stellar radii of the central star (much closer than H$_2$O masers are emitted), were not detected in these warm objects before. An exceptional case is the RV Tau variable, R Sct, with spectral type K0Ib. This object exhibits strong SiO and weak H$_2$O masers (I. Yamamura, 2004 private communication) as well as the 4 $\\mu$m SiO first overtone bands \\citep{mat02}. The RV Tau variables are believed to be low-mass post-AGB stars \\citep{jul86} with low metal abundances (population II; \\cite{gir00}), though these are spectroscopically classified as supergiants. Their spectral types change between K and M-type with light variation \\citep{pol97}. The atmosphere of late-type supergiants are not in hydrostatic equilibrium; effective temperature increases with decreasing metalicity \\citep{lev05}. The RV Tau variables are enshrouded by dust shell, and CO emission has been detected in two of these variables \\citep{buj88}. Although the optical counterpart of IRAS 22480+6002 is not reported to show any strong light variability (e.g., TASS; The Amateur Sky Survey\\footnote{data available at http://www.tass-survey.org/}), the optical spectroscopic classification, middle infrared properties, and maser characteristics of IRAS 22480+6002 indicate a close similarity to the properties of the RV Tau variables." }, "0710/0710.4525_arXiv.txt": { "abstract": "We report a sensitive search for the \\bhcn\\ emission line towards SDSS\\,J114816.64+525150.3 (hereafter:\\ J1148+5251) at $z$=6.42 with the Very Large Array (VLA). HCN emission is a star formation indicator, tracing dense molecular hydrogen gas ($n({\\rm H_2}) \\geq 10^4\\,$cm$^{-3}$) within star-forming molecular clouds. No emission was detected in the deep interferometer maps of J1148+5251. We derive a limit for the HCN line luminosity of $L'_{\\rm HCN} < 3.3 \\times 10^{9}\\,$K \\kms pc$^2$, corresponding to a HCN/CO luminosity ratio of $L'_{\\rm HCN}$/$L'_{\\rm CO}$$<$0.13. This limit is consistent with a fraction of dense molecular gas in J1148+5251 within the range of nearby ultraluminous infrared galaxies (ULIRGs; median value:\\ $L'_{\\rm HCN}$/$L'_{\\rm CO}$=0.17$^{+0.05}_{-0.08}$) and HCN-detected $z$$>$2 galaxies (0.17$^{+0.09}_{-0.08}$). The relationship between $L'_{\\rm HCN}$ and $L_{\\rm FIR}$ is considered to be a measure for the efficiency at which stars form out of dense gas. In the nearby universe, these quantities show a linear correlation, and thus, a practically constant average ratio. In J1148+5251, we find $L_{\\rm FIR}$/$L'_{\\rm HCN}$$>$6600. This is significantly higher than the average ratios for normal nearby spiral galaxies ($L_{\\rm FIR}$/$L'_{\\rm HCN}$=580$^{+510}_{-270}$) and ULIRGs (740$^{+505}_{-50}$), but consistent with a rising trend as indicated by other $z$$>$2 galaxies (predominantly quasars; 1525$^{+1300}_{-475}$). It is unlikely that this rising trend can be accounted for by a contribution of active galactic nucleus (AGN) heating to $L_{\\rm FIR}$ alone, and may hint at a higher median gas density and/or elevated star-formation efficiency toward the more luminous high-redshift systems. There is marginal evidence that the $L_{\\rm FIR}$/$L'_{\\rm HCN}$ ratio in J1148+5251 may even exceed the rising trend set by other $z$$>$2 galaxies; however, only future facilities with very large collecting areas such as the Square Kilometre Array (SKA) will offer the sensitivity required to further investigate this question. ", "introduction": "High redshift galaxy populations are now being detected back to 780 million years after the Big Bang (spectroscopically confirmed:\\ $z$=6.96; Iye \\etal\\ \\citeyear{iye06}), probing into the epoch of cosmic reionization (e.g., Fan \\etal\\ \\citeyear{fan06}; Hu \\& Cowie \\citeyear{hu06}). Many of these very distant galaxies show evidence for star formation activity (e.g., Taniguchi \\etal\\ \\citeyear{tan05}). Some are even found to be hyperluminous infrared galaxies (HLIRGs; Bertoldi \\etal\\ \\citeyear{ber03a}; Wang \\etal\\ \\citeyear{wan07}) with far-infrared (FIR) luminosities exceeding 10$^{13}$\\,\\lsol, suggesting vigorous star formation and/or AGN activity. To probe the earliest stages of galaxy formation and the importance of AGN in this process, it is necessary to study the star formation characteristics of these galaxies. A good diagnostic to examine the star-forming environments in distant HLIRGs are observations of molecular gas, the fuel for star formation. The by far brightest and most common indicator of molecular gas in galaxies is line emission from the rotational transitions of carbon monoxide (CO), which was detected in $\\sim$40 galaxies at high redshift ($z$$>$1; see Solomon \\& Vanden Bout \\citeyear{sv05} for a review). These observations have revealed molecular gas reservoirs with masses of $>$10$^{10}$\\,\\msol\\ in these galaxies, even in the highest redshift quasar known, J1148+5251 at $z$=6.42 (Walter \\etal\\ \\citeyear{wal03}, \\citeyear{wal04}; Bertoldi \\etal\\ \\citeyear{ber03b}). Although CO is a good tracer of the total amount of molecular gas in a galaxy, due to the relatively low critical density of $n_{\\rm H_2} \\sim 10^2-10^3\\,$cm$^{-3}$ required to collisionally excite its lower $J$ transitions, it is not a reliable tracer of the dense molecular cloud cores where the actual star formation takes place. Recent studies of nearby actively star-forming galaxies have shown that hydrogen cyanide (HCN) is a far better tracer of the dense ($n_{\\rm H_2} \\sim 10^5-10^6$\\,cm$^{-3}$) molecular gas where stars actually form (e.g.\\ Gao \\& Solomon \\citeyear{gao04a}, \\citeyear{gao04b}, hereafter:\\ GS04a, GS04b). In the local universe it was found that the HCN luminosity ($L'_{\\rm HCN}$) scales linearly (unlike $L'_{\\rm CO}$) with the FIR luminosity ($L_{\\rm FIR}$) over 7--8 orders of magnitude, ranging from Galactic dense cores to ULIRGs (Wu et al.\\ \\citeyear{wu05}). As $L_{\\rm FIR}$ traces the massive star formation rate (unless AGN heating is significant), this implies that HCN is also a good tracer of star formation. HCN has now also been detected in five galaxies at $z$$>$2 (Solomon \\etal\\ \\citeyear{sol03}; Vanden Bout \\etal\\ \\citeyear{vdb04}; Carilli \\etal\\ \\citeyear{car05}, hereafter:\\ C05; Wagg \\etal\\ \\citeyear{wag05}; Gao \\etal\\ \\citeyear{gao07}, hereafter:\\ G07). Adding a number of upper limits obtained for other high-$z$ galaxies, these observations indicate that the more luminous, higher redshift systems systematically deviate from the linear $L'_{\\rm HCN}$--$L_{\\rm FIR}$ correlation found in the local universe (G07), and hint at a rising slope of the relation toward high $L_{\\rm FIR}$ and/or $z$. To further investigate this apparent non-linear, rising trend, our aim has been to extend the range of existing HCN observations beyond redshift 6 and to higher $L_{\\rm FIR}$. In this letter, we report sensitive VLA\\footnote{The Very Large Array is a facility of the National Radio Astronomy Observatory, operated by Associated Universities, Inc., under cooperative agreement with the National Science Foundation.} observations of \\bhcn\\ emission toward the $z$=6.42 quasar J1148+5251, the highest redshift source detected in CO. A previous, less sensitive search for \\bhcn\\ emission in this source has yielded no detection (C05). We use a concordance, flat $\\Lambda$CDM cosmology throughout, with $H_0$=71\\,\\kms\\,Mpc$^{-1}$, $\\Omega_{\\rm M}$=0.27, and $\\Omega_{\\Lambda}$=0.73 (Spergel \\etal\\ \\citeyear{spe06}). ", "conclusions": "\\subsection{Median Gas Density and Star Formation Efficiency} Krumholz \\& Thompson (\\citeyear{kt07}) argue that $L_{\\rm FIR}$/$L'_{\\rm HCN}$ is expected to be higher for galaxies with a median (molecular) gas density $n_{\\rm med}$ close to or higher than the critical density $n_{\\rm crit}^{\\rm HCN}$ required for excitation of the observed HCN transition than for galaxies with lower $n_{\\rm med}$. In their case, they define star formation efficiency as the fraction of the mass that is converted into stars per dynamical time of the system. Note that this is different than the star formation rate per unit total gas mass. They argue that the non-linear relation between $L_{\\rm FIR}$ and $L'_{\\rm CO}$ (e.g., Kennicutt \\citeyear{ken98a}, \\citeyear{ken98b}; GS04b; Riechers \\etal\\ \\citeyear{rie06b}) arises due to the fact that CO traces all gas. The star-formation rate is then dictated by the density $n$ divided by the free-fall time $\\tau_{\\rm ff}$ ($\\tau_{\\rm ff} \\propto n^{-0.5}$), giving the standard Schmidt-law: star formation rate $\\propto n^{1.5}$, or $L_{\\rm FIR}$ $\\propto (L'_{\\rm CO})$$^{1.5}$. For molecules like HCN, which only trace the small fraction of dense gas clouds directly associated with star formation in normal galaxies, $\\tau_{\\rm ff}$ is roughly fixed by $n_{\\rm crit}$. Hence the star formation rate shows a linear relationship with $n$, or $L_{\\rm FIR}$ $\\propto (L'_{\\rm HCN})$$^{1.0}$. However, in extreme galaxies, where $n_{\\rm med}$ in the molecular ISM approaches $n_{\\rm crit}^{\\rm HCN}$, $\\tau_{\\rm ff}$ again becomes relevant (i.e., HCN emission no longer selects just the rare, dense peaks whose density is fixed by $n_{\\rm crit}^{\\rm HCN}$, but instead traces the bulk of the ISM, whose density can vary from galaxy to galaxy, and thus the variation of $n$ and $\\tau_{\\rm ff}$ re-enter the calculation), and the relationship approaches $L_{\\rm FIR}$ $\\propto (L'_{\\rm HCN})$$^{1.5}$ (and $L'_{\\rm HCN}$ $\\propto L'_{\\rm CO}$). Interestingly, current data show a marginal trend for a changing power-law index at the highest luminosities of the type proposed by Krumholz \\& Thompson. This change in power-law index from 1 to 1.5 would suggest that, in these extreme luminosity systems, $n_{\\rm med}$ approaches $n_{\\rm crit}^{\\rm HCN}$. More systems at high luminosity are required to confirm this trend of changing power-law index. \\subsection{The Role of AGN Heating for $L_{\\rm FIR}$} Like most of the $z$$>$2 HCN-detected sources, J1148+5251 is a quasar. It has been found that, even for such strong AGN galaxies, the bulk of $L_{\\rm FIR}$ is likely dominatly heated by star formation in most cases (e.g., Carilli \\etal\\ \\citeyear{car01}; Omont \\etal\\ \\citeyear{omo01}; Beelen \\etal\\ \\citeyear{bee06}; Riechers \\etal\\ \\citeyear{rie06b}). However, based on radiative transfer models of the dust SED of J1148+5251, Li \\etal\\ (\\citeyear{li07}) argue that this source may currently undergo a `quasar phase', in which AGN heating of the hot and warm dust contributes significantly to $L_{\\rm FIR}$. If correct, this may be an alternative explanation for the elevated $L_{\\rm FIR}$/$L'_{\\rm HCN}$ in this galaxy. The (rest-frame) IR properties (tracing emission from hot dust) of J1148+5251 are similar to those of other $z$$>$6 quasars with much lower $L_{\\rm FIR}$ (tracing emission from warm dust), and even to local quasars (Jiang \\etal\\ \\citeyear{jia06}). This supports the assumption that the hot dust in J1148+5251 is dominantly heated by the AGN; however, the lack of a correlation between $L_{\\rm IR}$ and $L_{\\rm FIR}$ in quasars indicates that the warm dust may still be dominantly heated by star formation. Moreover, J1148+5251 follows the radio-FIR correlation for star-forming galaxies (Carilli \\etal\\ \\citeyear{car04}), which also suggests a starburst origin for the dominant fraction of $L_{\\rm FIR}$. Furthermore, one of the $z$$>$2 HCN detections and some of the meaningful limits are submillimeter galaxies without a known luminous AGN, but are still offset from the local $L_{\\rm FIR}$/$L'_{\\rm HCN}$ relation. It thus appears unlikely that AGN heating alone can account for the higher average $L_{\\rm FIR}$/$L'_{\\rm HCN}$ in the high-$z$ galaxy sample. \\subsection{Implications for Future Studies} Even when assuming the highest $L_{\\rm FIR}$/$L'_{\\rm HCN}$ of 2835 found among all HCN-detected galaxies in Table \\ref{tab-2}, the depth of our observations is sufficient to detect a galaxy with the redshift and $L_{\\rm FIR}$ of J1148+5251 ($z$=6.42) in HCN emission at a signal-to-noise ratio of $>$4.5. To first order, our lower limit thus is consistent with previous suggestions (G07) that $L_{\\rm FIR}$/$L'_{\\rm HCN}$ ratios in high redshift sources lie systematically above those for nearby galaxies. The scatter around this trend is still significant, and will primarily be improved by increasing the number of HCN-detected galaxies at high $z$. In addition, it will be important to improve on the main sources of error for the individual high-$z$ detections (e.g., signal-to-noise limited HCN/CO linewidth ratio, accuracy of the FIR SED fit, AGN bias of $L_{\\rm FIR}$). The statistical and individual results, so far, would even be consistent with an even stronger increase in $L_{\\rm FIR}$/$L'_{\\rm HCN}$ toward the highest $z$ and/or $L_{\\rm FIR}$. Our study of J1148+5251 may hint at such an effect. Clearly, it is desirable to obtain more sensitive observations of this source to further investigate this issue. Due to its superior collecting area and high calibrational stability, the VLA is ideally suited for such a sensitive study. Although J1148+5251 is the most CO- and FIR-luminous $z$$>$6 galaxy known, 80\\,hr of VLA observations were necessary to obtain the current limit. In a favourable case, the \\bhcn\\ line may have a strength of about 1.5 times the current rms. To obtain a solid 5\\,$\\sigma$ detection of such a line, of order 1000\\,hr of observations with the VLA would be required. Due to improved receivers and antenna performance, the fully operational EVLA will be by a factor of two more sensitive to spectral lines of several 100\\,\\kms\\ width (such as in J1148+5251), but will still require long integration times. Studies of dense gas at $z$$>$6 thus appear to require an order of magnitude increase in collecting area, such as offered by future facilities like the SKA phase I demonstrator (e.g., Carilli \\citeyear{car06}), which can serve as a low frequency counterpart to the Atacama Large Millimeter/submillimeter Array (ALMA)." }, "0710/0710.4149_arXiv.txt": { "abstract": "{The Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE) is an atmospheric Cherenkov telescope (ACT) that uses a large mirror array to achieve a relatively low energy threshold. For sources with Crab-like spectra, at high elevations, the detector response peaks near 100 GeV. Gamma-ray burst (GRB) observations have been a high priority for the STACEE collaboration since the inception of the experiment. We present the results of 20 GRB follow-up observations at times ranging from 3 minutes to 15 hours after the burst triggers. Where redshift measurements are available, we place constraints on the intrinsic high-energy spectra of the bursts.} \\begin{document} ", "introduction": "The Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE) is a showerfront-sampling Cherenkov telescope sensitive to gamma rays above 100 GeV. It is located at the National Solar Thermal Test Facility (NSTTF) at Sandia National Laboratories outside Albuquerque, New Mexico, USA. The NSTTF is located at 34.96$^{\\circ}$N, 106.51$^{\\circ}$W and is 1700 m above sea level. The facility has 220 heliostat mirrors designed to track the sun across the sky, each with 37 m$^{2}$ area. STACEE uses 64 of these heliostats to collect Cherenkov light produced by air showers. STACEE employs five secondary mirrors on the solar tower to focus the Cherenkov light onto photomultiplier tube (PMT) cameras, as shown in Figure \\ref{concept}. The light from each heliostat is detected by a separate PMT and the waveform of the PMT signal is recorded by a flash ADC. A programmable digital delay and trigger system\\cite{IEEENSS2000} selects showers for acquisition while eliminating most random coincidences of night sky background photons. Under good observing conditions, STACEE operates with a threshold around 5 photoelectrons per channel. A detailed description of the instrument can be found in D.M. Gingrich et al.\\cite{Gingrich05}. \\begin{figure}[t] \\begin{minipage}{0.5\\textwidth} \\begin{center} \\includegraphics[width=.9\\textwidth]{concept5.eps} \\end{center} \\end{minipage} \\hfill \\begin{minipage}{0.45\\textwidth} \\caption{Conceptual drawing of STACEE.} \\label{concept} \\vskip 1in \\end{minipage} \\end{figure} The large mirror area of the STACEE detector leads to an energy threshold lower than those attainable by most single-dish imaging telescopes or water-Cherenkov telescopes. The energy threshold - defined as the energy at which the trigger rate peaks - is determined by the spectrum of the source and the effective area of the detector at the target position. For targets above 60$^{\\circ}$ in elevation with power-law spectral indices between 2 and 3, the energy threshold is typically between 150 and 200 GeV. For targets near zenith, STACEE has significant effective area at energies as low as 50 GeV. A low energy threshold opens up the possibility of detecting more distant sources\\cite{Primack05}. Collisions of gamma rays with extragalactic background light (EBL) photons produce electron-positron pairs, attenuating the gamma-ray flux from distant sources. The extinction becomes more severe with increasing energy, producing an energy-dependent horizon for gamma-ray observations. Thus, a low energy threshold is advantageous when attempting to characterize the high-energy emission of GRBs, for which the mean measured redshift (for Swift bursts) is 2.7\\cite{Le07}. STACEE observations are typically taken in pairs of on-source and off-source runs. The off-source runs serve as measurements of the background event rate produced by cosmic-ray showers. Under normal conditions, the cosmic-ray trigger rate is $\\sim$5 Hz. Background rejection techniques have been explored and the most effective technique is described elsewhere\\cite{Kildea05,Kildea05_2,Lindner06}. After background rejection cuts, STACEE typically obtains a 5$\\sigma$ detection of the Crab with approximately 10 hours of on-source observations. Under good observing conditions, STACEE would obtain a 5$\\sigma$ detection in 30 minutes for a source with a spectrum equal to 4.5 times that of the Crab. ", "conclusions": "" }, "0710/0710.1841_arXiv.txt": { "abstract": "We report the discovery of a planet transiting a moderately bright ($V=12.00$) G star, with an orbital period of $2.788491\\pm0.000025$ days. From the transit light curve we determine that the radius of the planet is $\\rpl = 1.257\\pm0.053\\,\\rjuplong$. \\hatcurb\\ has a mass of $\\mpl = \\hatcurm\\,\\mjuplong$, similar to the average mass of previously-known transiting exoplanets, and a density of $\\rhopl = \\hatcurrho\\,\\gcmc$. We find that the center of transit is $T_{\\mathrm{c}} = \\hatcurT$ (HJD), and the total transit duration is \\hatcurdur\\,days. ", "introduction": "\\label{sec:intro} To date about 20 extrasolar planets have been found which transit their parent stars and thus yield values for their mass and radius\\footnote{Extrasolar Planets Encyclopedia: http://exoplanet.eu}. Masses range from 0.07\\,\\mjup\\ \\citep[GJ436;][]{gillon07} to about 9\\,\\mjup\\ \\citep[HAT-P-2b;][]{bakos07}, and radii from 0.7\\,\\rjup\\ (GJ436) to about 1.7\\,\\rjup\\ \\citep[TRES-4;][]{mandushev07}. These data provide an opportunity to compare observations with theoretical models of planetary structure across a wide range of parameters, including those of the host star \\citep[e.g.][and references therein]{burrows07,fortney07}. Transits also yield precise determination of other physical parameters of the extrasolar planets, for instance the surface gravity. Interesting correlations between these parameters have been noted early on, such as that between masses and periods \\citep{mazeh05} or periods and surface gravities \\citep{southworth07}. Classes of these close-in planets have also been suggested, such as very hot Jupiters (VHJs; P=1--3 days) and hot Jupiters \\citep[HJs; P=3--9 days;][]{gaudi05}, or a possible dichotomy based on Safronov numbers \\citep{hansen07}. However, the small ensemble of transiting exoplanets (TEPs) does not allow robust conclusions, thus the addition of new discoveries is valuable. Over the past year the HATNet project\\footnote{www.hatnet.hu} \\citep{bakos02,bakos04}, a wide-angle photometric survey, has announced four TEPs. In this Letter we report on the detection of a new transiting exoplanet, which we label \\hatcurb, and our determination of its parameters, such as mass, radius, density and surface gravity. ", "conclusions": "\\label{sec:disc} \\hatcurb\\ is an ordinary hot Jupiter (P = 2.788 days) with slightly inflated radius (\\rpl = 1.26\\rjup) for its mass of 1.06\\mjup, orbiting a slightly metal rich solar-like star. The $\\sim$20\\% radius inflation is what current models predict for a planet with equilibrium temperature of $\\sim$1500K \\citep{burrows07,fortney07}. \\hatcurb\\ is more massive than any of the known TEPs with similar period ($2.5\\lesssim P \\lesssim3$\\,d), such as XO-2b, WASP-1, HAT-P-3b, TRES-1, and HAT-P-4b, with the exception of TRES-2. The latter is fairly similar in mass, radius, orbital period, and stellar effective temperature. However, \\hatcurb\\ is interesting in that it falls between Class I and II, as defined by the Safronov number and $T_{eq}$ of the planet \\citep{hansen07}. \\hatcurb\\ has a Safronov number of $0.059\\pm0.005$ , while Class I is defined as $0.070\\pm0.01$, especially at $T_{eq}\\sim 1500$K. It seems that the additional discovery and characterization of transiting planets of Jupiter and higher masses would be very helpful in order to understand these new correlations and their reality." }, "0710/0710.1246_arXiv.txt": { "abstract": "We present an algorithm for solving the linear dispersion relation in an inhomogeneous, magnetised, relativistic plasma. The method is a generalisation of a previously reported algorithm that was limited to the homogeneous case. The extension involves projecting the spatial dependence of the perturbations onto a set of basis functions that satisfy the boundary conditions (spectral Galerkin method). To test this algorithm in the homogeneous case, we derive an analytical expression for the growth rate of the Weibel instability for a relativistic Maxwellian distribution and compare it with the numerical results. In the inhomogeneous case, we present solutions of the dispersion relation for the relativistic tearing mode, making no assumption about the thickness of the current sheet, and check the numerical method against the analytical expression. ", "introduction": "Dissipation of the energy carried by relativistic plasma outflows is important for the physics of pulsar winds and gamma-ray bursts (for recent reviews see \\cite{2007astro.ph..3116K} and \\cite{2005RvMP...76.1143P}). In these objects, the plasma is probably composed of electrons, positrons and protons. As well as being in relativistic bulk motion with respect to the observer, the random thermal energy of the plasma may also be relativistic, i.e., comparable to the rest mass energy of the constituent particles. In this paper we concentrate on two instabilities: these are the two-stream or Weibel instability~\\cite{1959PhRvL...2...83W} and the tearing modes in a relativistic neutral pair plasma current sheet, thought to play a role in the formation process of relativistic shocks \\cite{2007arXiv0706.3126S} and the dissipation of magnetic energy in pulsar winds \\cite{2005PPCF...47B.719K}. They are investigated by generalising and extending to the inhomogeneous case the method presented in \\cite{2007PPCF...49..297P}. Motivated primarily by the need for code verification, we have derived some analytical expressions for the linear growth rates of these instabilities. The Weibel instability is very important in astrophysical processes because it is able to generate a magnetic field by extracting the free energy from an anisotropic momentum distribution in an unmagnetised plasma or from the kinetic drift energy. There is an extensive literature on the Weibel instability: general conditions for the existence of the relativistic Weibel instability for arbitrary distribution functions are discussed in~\\cite{2006PhPl...13b2107S}, and wave propagation in counter-streaming magnetised nonrelativistic Maxwellian plasmas are studied in~\\cite{2005PhPl...12.2901T,2006PhPl...13f2901T}. Dispersion curves have been found in some special cases such as, for example, the fully relativistic bi-Maxwellian distribution function, (Yoon~\\cite{1989PhFlB...1.1336Y}), and the water-bag distribution function, in which case closed-form analytical expressions can be derived not only for the Weibel instability~(Yoon~\\cite{1987PhRvA..35.2718Y}), but also for the cyclotron maser and whistler instabilities (Yoon~\\cite{1987PhRvA..35.2619Y}). However, finding an analytical expression for the dispersion relation for a given equilibrium distribution function is a complicated or even impossible task. It involves a four-dimensional integration (3D in momentum space and 1D in time) of the equilibrium distribution function which is difficult to perform in closed form. For this reason, the water-bag distribution is the preferred profile to analyse magnetic field generation in fast ignitor scenarios, (Silva et al.~\\cite{2002PhPl....9.2458S}) and in relativistic shocks, (Wiersma and Achterberg~\\cite{2004A&A...428..365W}, Lyubarsky and Eichler~\\cite{2006ApJ...647.1250L}). The Weibel instability in a magnetised electron-positron pair plasma has been investigated by Yang et al~\\cite{1993PhFlB...5.3369Y} using two model distributions: the water bag, and one with a power-law dependence at high energy. A general covariant description has been formulated by Melrose~\\cite{1982AuJPh..35...41M} and by Schlickeiser~\\cite{2004PhPl...11g5532S}. In the present work, we focus on equilibrium configurations given by a relativistic Maxwellian distribution function, which allows one to reduce the four-dimensional integral to a simple one-dimensional integral, as we demonstrate in Section~\\ref{sec:weibel_analytical}. The growth rates are then found by solving this equation using a single numerical quadrature, and are compared to the results found using our extended algorithm in Section~\\ref{sec:weibel_numerical}. In the inhomogeneous case, the stability properties of a nonrelativistic Harris current sheet also have a substantial literature, with notable recent studies by Daughton~\\cite{1999PhPl....6.1329D} and Silin et al.~\\cite{2002PhPl....9.1104S}. In the relativistic case, the tearing mode instability has been investigated by Zelenyi \\& Krasnoselskikh~\\cite{1979SvA....23..460Z}, by integrating first order perturbations of the relativistic Maxwellian distribution function along approximate, straight-line particle trajectories, in the thick layer limit (in which the Larmor radius is much smaller than the thickness of the current sheet). In Section~\\ref{sec:tearing_analytical} we lift these restrictions to present new results for the tearing mode instability in a neutral current sheet of arbitrary temperature and thickness, and compare these with the results found using the generalised algorithm in Section~\\ref{sec:tearing_numerical}. In this work, no assumption is made about the thickness of the current sheet, and the particle trajectories are found numerically in the background magnetic field. The numerical method, which is an extension of our previous algorithm, \\cite{2007PPCF...49..297P}, that computes the linear dispersion relation of waves within a Vlasov-Maxwell description, is described in Section~\\ref{sec:algorithm}. It is based on the approach of Daughton~\\cite{1999PhPl....6.1329D} for non relativistic Maxwellians, and involves explicit time integration of particle orbits along the unperturbed trajectories. We modify and extend our former code to include inhomogeneities in the plasma equilibrium configuration. Moreover, we generalise it to a fully relativistic approach, i.e., one that allows for relativistic temperatures as well as relativistic drift speeds. ", "conclusions": "\\label{sec:conclusion} We present a generalisation and extension of our previous algorithm~\\cite{2007PPCF...49..297P} to solve the linear dispersion relation for relativistic multi-component inhomogeneous and magnetised plasmas. The code is validated by comparing the results with two standard configurations: the relativistic Weibel instability in a homogeneous plasma, and the tearing mode instability in a relativistic neutral Harris sheet. To effect the comparison, we derived useful analytical expressions, Eq.~(\\ref{eq:RelDispWeibel}) and Eq.~(\\ref{eq:Harris2}), for the dispersion relations in these configurations and solved them numerically. We conclude that this code is a suitable tool for the study of stability properties of more general configurations of interest in gamma-ray burst and pulsar wind theories. \\ack{This research was supported by a grant from the G.I.F., the German-Israeli Foundation for Scientific Research and Development.}" }, "0710/0710.3769_arXiv.txt": { "abstract": "{We describe a broad class of time-dependent exact wave solutions to 6D gauged chiral supergravity with two compact dimensions. These 6D solutions are nontrivial warped generalizations of 4D pp-waves and Kundt class solutions and describe how a broad class of previously-static compactifications from 6D to 4D (sourced by two 3-branes) respond to waves moving along one of the uncompactified directions. Because our methods are generally applicable to any higher dimensional supergravity they are likely to be of use for finding the supergravity limit of time-dependent solutions in string theory. The 6D solutions are interesting in their own right, describing 6D shock waves induced by high energy particles on the branes, and as descriptions of the near-brane limit of the transient wavefront arising from a local bubble-nucleation event on one of the branes, such as might occur if a tension-changing phase transition were to occur.} \\preprint{PI-COSMO-65} \\begin{document} ", "introduction": "Understanding time-dependent dynamics is central to applications of higher-dimensional supergravity theories \\cite{HiDSugra} to cosmology and to particle physics. Inasmuch as higher-dimensional supergravities provide the low-energy limit of string theories, any understanding of time-dependence in the supergravity limit also provides a guide for the thornier issue of understanding these same issues in string theory. For these reasons there is considerable interest in finding time-dependent solutions to higher-dimensional supergravity \\cite{tdepsusy} (as well as of non-supersymmetric gravity \\cite{tdepnonsusy}, since this can also sometimes capture similar physics). Much of this activity has focussed on 10D and 5D theories (motivated by string applications, and Randall-Sundrum \\cite{RS} constructions), although these can either be more difficult to solve or they can have features which are specific to the relative simplicity of co-dimension one spaces. Six-dimensional supergravity has more recently emerged as being a useful intermediate workshop within which to investigate phenomena which can often generalize to still higher-dimensional contexts. Interest in 6D supergravity has been further sharpened by the recognition that it can provide insights into the nature of the cosmological constant problem \\cite{SLED1}--\\cite{SLEDpheno}, by building on the observation that higher-dimensional theories can break the link between the 4D vacuum energy density and the curvature of 4D spacetime \\cite{5DSelfTune}--\\cite{6DNonSUSYSelfTunex} (see \\cite{Burgess:2007ui} for a review). There has also been considerable recent interest in 6 dimensional models more generally \\cite{Closelyrelatedworks}. Including branes is notoriously difficult due the necessity to regularize UV divergences which arise \\cite{regularizations}, however for recent work on understanding this more deeply in the context of effective field theory see \\cite{claudiaeft}. In this paper we further the program of understanding time-dependent solutions to higher-dimensional supergravity in two ways. In \\S2, we present a method for constructing explicit exact solutions to the supergravity field equations of 6D gauged chiral supergravity, to derive a new class of exact solutions to these equations which describe a class of gravitational waves passing through spacetimes for which two dimensions are compactified in response to the stress-energy of two space-filling 3-branes. In \\S3 we discuss the applications of these solutions, including how to construct the shock wave metric corresponding to ultra-relativistic particles and BH's moving on one of the branes, and also how these solutions provide some insight into the transient part of the dynamics describing outgoing waves which would arise shortly after a phase transition on one of the source branes. ", "conclusions": "Six dimensional supergravity provides a fruitful laboratory for investigating the issues which underly higher-dimensional physics in general, and brane approaches to the cosmological constant problem in particular. It does so because 6D is rich enough to exhibit many of the properties of still-higher dimensions --- like moduli-stabilization through fluxes \\cite{Susha}, brane back-reaction on internal geometries, chiral fermions and Green-Schwarz anomaly cancellation \\cite{6Danomalies}. Yet it is also on the one hand simple enough to allow the development of techniques for obtaining physically interesting exact solutions, but on the other hand not so simple as to be misleading about what happens in higher dimensional (in a way which co-dimension one physics sometimes can be). We have used these properties to explore solution-generation techniques which we believe to be applicable to a wide variety of higher-dimensional supergravities. We do so by using these techniques to construct a new class of time-dependent exact solutions to the field equations of 6D chiral gauged supergravity. These solutions describe the physics of nonlinear gravitational waves passing through compactified spacetimes for which two dimensions are self-consistently compactified in response to the presence of two space-filling branes and a bulk Maxwell flux. We believe these methods to merit more detailed exploration." }, "0710/0710.0901_arXiv.txt": { "abstract": "The diffuse interstellar bands (DIBs) probably arise from complex organic molecules whose strength in local galaxies correlates with neutral hydrogen column density, \\nhi, and dust reddening, \\ebmv. Since Ca{\\sc \\,ii} absorbers in quasar (QSO) spectra are posited to have high \\nhi\\ and significant \\ebmv, they represent promising sites for the detection of DIBs at cosmological distances. Here we present the results from the first search for DIBs in 9 Ca{\\sc \\,ii}-selected absorbers at $0.07 < z_{\\rm abs} < 0.55$. We detect the 5780\\,\\AA\\ DIB in one line of sight at $z_{\\rm abs} = 0.1556$; this is only the second QSO absorber in which a DIB has been detected. Unlike the majority of local DIB sight-lines, both QSO absorbers with detected DIBs show weak 6284\\,\\AA\\ absorption compared with the 5780\\,\\AA\\ band. This may be indicative of different physical conditions in intermediate redshift QSO absorbers compared with local galaxies. Assuming that local relations between the 5780\\,\\AA\\ DIB strength and \\nhi\\ and \\ebmv\\ apply in QSO absorbers, DIB detections and limits can be used to derive \\nhi\\ and \\ebmv. For the one absorber in this study with a detected DIB, we derive \\ebmv\\ = 0.23\\,mag and $\\log$\\nhi\\ $\\ge$ 20.9, consistent with previous conclusions that Ca{\\sc \\,ii} systems have high H{\\sc \\,i} column densities and significant reddening. For the remaining 8 Ca{\\sc \\,ii}-selected absorbers with 5780\\,\\AA\\ DIB non-detections, we derive \\ebmv\\ upper limits of 0.1--0.3\\,mag. ", "introduction": "Damped Lyman alpha (DLA) systems are usually considered to be the class of QSO absorber with the highest neutral hydrogen column densities [\\nhi\\ $\\ge 2 \\times 10^{20}$ \\cm]. Nonetheless, the DLAs are characterised by generally low metallicities and gas-phase depletion fractions \\citep[e.g.][]{KhareP_04a,AkermanC_05a,ProchaskaJ_07a} and low reddening due to dust (\\citealt{MurphyM_04c}; \\citealt*{EllisonS_05a}). The DLAs are also poor in molecules, as demonstrated by both their generally low fractions of H$_2$ \\citep[e.g.][]{LedouxC_03a} and the lack of a detection for any other molecular species, such as OH or CO \\citep[e.g.][]{CurranS_06a}. Although the handful of DLAs which do exhibit molecular H$_2$ absorption may be biased, e.g. towards high metallicities \\citep{PetitjeanP_06a}, such systems can offer a novel insight into the physical conditions of the galactic interstellar medium \\citep[ISM, e.g.][]{SrianandR_05a,NoterdaemeP_07a}. In addition to the study of H$_2$, one avenue that is just starting to be explored is how the diffuse interstellar bands \\citep[DIBs; see reviews by][]{HerbigG_95a,SarreP_06a} may be used to probe the intermediate redshift ISM. Although lacking definitive identifications, the strength (both absolute and relative) of these broad absorption features in the Milky Way (MW) and other nearby galaxies exhibit dependencies on (and sometimes, tight correlations with) neutral gas content, dust reddening, metallicity and local radiation field \\citep[e.g.][]{HerbigG_93a,CoxN_06a,WeltyD_06a,CoxN_07a}. Moreover, if DIBs are as strong in DLAs as they are in the MW [i.e.~for a given \\nhi], then they should be relatively easy to detect at intermediate redshifts. The first systematic search for DIBs in DLAs has recently been carried out by Lawton et al.~(in preparation) in 7 $z < 1$ absorbers. In only one case were DIBs detected: the 4428, 5705 and 5780\\,\\AA\\ features\\footnote{We cite all DIBs with reference to their normal air wavelengths, although their vacuum values have been used in practice in order to be consistent with our spectral wavelength calibration; see Section 2.} were all detected in the $z \\sim 0.5$ DLA towards AO 0235$+$164 \\citep{JunkkarinenV_04a,YorkB_06a}. Lawton et al.~showed that for the 6 non-detections in their DLA sample, the strength of the 5780\\,\\AA\\ DIB [which shows one of the tightest correlations with \\nhi\\ in the MW] is often at least 3 times weaker in DLAs for a given \\nhi\\ compared with Galactic sight-lines. The 6284\\,\\AA\\ DIB is even more under-abundant in DLAs for a given \\nhi: 4--10 times weaker than towards Galactic sight-lines. A similar result has been found for DIBs in the Large and Small Magellanic Cloud \\citep[LMC and SMC;][]{WeltyD_06a} where the 5780\\,\\AA\\ DIB is typically 10--30 times weaker than expected from the Galactic relation. On the other hand, the 5780\\,\\AA\\ DIB strength correlates well with \\ebmv\\ in both Galactic and Magellanic Cloud sight-lines, and the detection towards AO 0235$+$164 also lies on the same relationship \\citep{YorkB_06a}. These results hint that DIB formation/survival and high dust content are closely linked and that DIBs are therefore most likely to be detected in galaxies with high reddening. \\citet*{WildV_06a} have recently suggested that absorbers identified via high equivalent widths (EWs) of Ca{\\sc \\,ii} may select the highest \\nhi\\ and highest \\ebmv\\ absorbers. For example, whereas DLAs have been constrained to have \\ebmv\\ $\\le$ 0.04 \\citep{MurphyM_04c,EllisonS_05a}, \\citet{WildV_06a} find that absorbers with Ca{\\sc \\,ii} $\\lambda$3934 EWs $>$0.7\\,\\AA\\ have \\ebmv\\ values up to $\\sim$0.1 mag. Ca{\\sc \\,ii} absorbers may therefore be promising sites for the detection of DIBs. ", "conclusions": "\\begin{figure} \\centerline{\\rotatebox{0}{\\resizebox{9cm}{!} {\\includegraphics{EBMV_NaI_5780.ps}}}} \\caption{\\label{ebmv_fig} Correlations of Na{\\sc \\,i} (top panel), \\nhi\\ (middle panel) and \\ebmv\\ (bottom panel) versus the log of the EW (in m\\AA) of the 5780\\,\\AA\\ DIB. Open squares are Galactic points from \\citet{HerbigG_93a}; open triangles/diamonds are SMC/LMC points from \\citet{WeltyD_06a}, \\citet{VladiloG_87a} and \\citet{Vidal-MadjarA_87a}; crosses are other extra-galactic data points taken from \\citet{SollermanJ_05a}, \\citet{DOdoricoS_89a} and \\citet{HeckmanT_00b}; solid stars are DLAs from \\citet{YorkB_06a} and Lawton et al.~(in preparation). The best (least squares) fit to the data in each panel is shown with a solid line and the fit solution given at the bottom of each panel. The fit of the 5780\\,\\AA\\ DIB with \\ebmv\\ uses all available data; the fits with of 5780 with \\nhi\\ and Na{\\sc \\,i} use just the Galactic data. In the middle and lower panels, the solid vertical line indicates the 5780\\,\\AA\\ DIB detection towards J0013$-$0024 and the dotted lines are the 3$\\sigma$ upper limits for 6 other sight-lines where we have a 5780\\,\\AA\\ limits (as given in Table \\ref{dib_table}). } \\end{figure} In local (e.g.~Galactic, LMC, SMC) sight-lines, the 6284\\,\\AA\\ DIB is typically 2--3 times stronger than the 5780\\,\\AA\\ DIB (e.g., York et al. 2006a and references therein). The one exception is the unusual SMC wing sight-line towards Sk~143 where the 6284\\,\\AA\\ DIB has an EW less than half that of the 5780\\,\\AA\\ DIB \\citep{WeltyD_06a}. \\citet{YorkB_06a} also found that in the one DLA sight-line with a 5780\\,\\AA\\ band detection out of the 7 studied by Lawton et al.~(in preparation), the EW of the 6284\\,\\AA\\ feature was also constrained to be less than the EW of the 5780\\,\\AA\\ line. \\citet{YorkB_06a} suggested that these unusual line ratios could be an indication of an ISM that is more protected from the ambient UV radiation field. In the Ca{\\sc \\,ii}-selected absorber towards J0013$-$0024, we constrain the EW of the 6284\\,\\AA\\ DIB to be at least $\\sim$20\\% weaker than the 5780\\,\\AA\\ band. The DIB ratios in this absorber are therefore consistent with those in the DLA detection of \\citet{YorkB_06a} and the SMC wing sight-line Sk~143 but inconsistent with other local sight-lines, including starburst galaxies \\citep{HeckmanT_00b} and the Magellanic Clouds \\citep{WeltyD_06a}. In the Galaxy, many DIBs show correlations with \\nhi\\ and $N$(Na{\\sc \\,i}) \\citep[e.g.][]{HerbigG_93a,HerbigG_95a}. In Table \\ref{dib_table} we tabulate the EWs of the Na{\\sc \\,i} doublet for our Ca{\\sc \\,ii}-selected absorbers. However, we do not calculate the Na{\\sc \\,i} column density because, if the lines are strong enough to be detected in our moderate resolution spectra, they are likely to be saturated. The Galactic correlations of \\nhi\\ and $N$(Na{\\sc \\,i}) with the 5780\\,\\AA\\ DIB, which is one of the tightest of the DIB relations, is shown in Fig. \\ref{ebmv_fig}. We also show data for the Magellanic Clouds \\citep{WeltyD_06a} and DLAs (\\citealt{YorkB_06a}; Lawton et al.~in preparation), where it can be seen that the DIBs are weak for their \\nhi\\ compared with the Galactic correlation. As shown in Fig. \\ref{ebmv_fig}, the DIBs in extra-galactic sight-lines are also weak for their Na{\\sc \\,i} column densities. These departures from the Galactic relations are probably due to a combination of effects including ambient radiation field, metallicity and dust-to-gas ratios \\citep{CoxN_06a}. Assuming that the Galactic 5780--\\nhi\\ relation provides a lower limit for the H{\\sc \\,i} column density, DIB detections may be useful for constraining \\nhi\\ in the absence of \\lya\\ observations. For example, \\citet{WildV_05a} and \\citet{WildV_06a} have argued that Ca{\\sc \\,ii} absorbers represent the high column density end of the DLA distribution. Our detection of the 5780\\,\\AA\\ DIB in the $z_{\\rm abs} = 0.1556$ absorber towards J0013$-$0024 supports this hypothesis, and we derive $\\log$\\nhi\\ $\\ge$ 20.9 for this absorber. Unlike correlations with \\nhi\\ and $N$(Na{\\sc \\,i}), \\citet{WeltyD_06a} have shown that the 5780\\,\\AA\\ DIB strength follows a single relationship with \\ebmv\\ in both Galactic and Magellanic Cloud sight-lines. \\citet{YorkB_06a} found that the single DLA 5780\\,\\AA\\ DIB detection towards AO 0235$+$164 fell on the same relationship. It is not yet clear whether the apparent universality of this correlation is driven by a tight physical connection between dust properties and DIB formation \\citep{CoxN_07a} or whether it is coincidence of different physical drivers working in different directions \\citep{CoxN_06a}. However, if the 5780--\\ebmv\\ is applicable to QSO absorbers, we can use our DIB detection limits to constrain their reddening. \\citet{WeltyD_06a} derive a best fit correlation between the 5780\\,\\AA\\ DIB (in m\\AA) and the \\ebmv\\ for Galactic sight-lines: log \\ebmv\\ = $-$2.70 + 1.01 log EW(5780). We derive the best fit relation to the 5780--\\ebmv\\ data points of the Galactic plus Magellanic Cloud plus AO 0235$+$164 DLA sight-lines and find log \\ebmv\\ = $-$2.19 + 0.79 log EW(5780) (see Figure \\ref{ebmv_fig}). The range in log \\ebmv\\ values around the best fit relation is $\\sim \\pm$ 0.4 dex. This correlation gives a reddening for the Ca{\\sc \\,ii} absorber towards J0013$-$0024 of \\ebmv$\\,\\,\\sim0.23$\\,mag and upper limits for the other 8 Ca{\\sc \\,ii} absorbers in our sample of 0.1--0.3\\,mag. These values provide independent estimates of reddening associated with Ca{\\sc \\,ii}-selected absorbers that do not depend directly on the choice of extinction law and can be applied for individual absorbers and not just in a statistical fashion \\citep[e.g.][]{MurphyM_04c,WildV_05a,WildV_06a}. The Ca{\\sc \\,ii} EWs of our sample are typically $<0.7$\\,\\AA\\ (see Table \\ref{dib_table}); for this range of EWs, \\citet{WildV_06a} determine average reddenings of \\ebmv\\ = 0.02, 0.03 and 0.03\\,mag for MW, LMC and SMC extinction curves respectively." }, "0710/0710.5523_arXiv.txt": { "abstract": "We have performed a high mass and force resolution simulation of an idealized galaxy forming from dissipational collapse of gas embedded in a spherical dark matter halo. The simulation includes star formation and effects of stellar feedback. In our simulation a stellar disk forms with a surface density profile consisting of an inner exponential breaking to a steeper outer exponential. The break forms early on and persists throughout the evolution, moving outward as more gas is able to cool and add mass to the disk. The parameters of the break are in excellent agreement with observations. The break corresponds with a rapid drop in the star formation rate associated with a drop in the cooled gas surface density, but the outer exponential is populated by stars that were scattered outward on nearly circular orbits from the inner disk by spiral arms. The resulting profile and its associated break are therefore a consequence of the interplay between a radial star formation cutoff and redistribution of stellar mass by secular processes. A consequence of such evolution is a sharp change in the radial mean stellar age profile at the break radius. ", "introduction": "\\label{sec:intro} Since the early work of \\citet{de-Vaucouleurs:1958} it has been recognized that the disks of spiral galaxies generally follow an exponential radial surface brightness profile, and various theories have explored the physical causes and consequences of this property \\citep[e.g.][]{Fall:1980lr, Lin:1987pb, Dalcanton:1997bh, Mo:1998mi, van-den-Bosch:2001aa}. However, since \\citet{van-der-Kruit:1979gb, van-der-Kruit:1987fk} it has been known that the outer regions of disks exhibit more varied behavior. This has been confirmed by an abundance of recent data \\citep[e.g.,][hereafter PT06]{Pohlen:2000ff, Pohlen:2002ec, Bland-Hawthorn:2005ys, Erwin:2005kl, Pohlen:2006lh}. In a sample of nearby late-type galaxies from the Sloan Digital Sky Survey, PT06 found that about 60\\% have an inner exponential followed by a steeper outer exponential (downward-bending), $\\sim$30\\% have the inner exponential followed by a shallower outer exponential (upward-bending), while only $\\sim10$\\% have no detectable breaks. Therefore breaks are a common feature of disk galaxies that any complete theory of galaxy formation must be able to explain. Furthermore, the discovery of UV emission at radii well beyond the H$\\alpha$ emission cutoff \\citep{Gil-de-Paz:2005, Thilker:2005, Thilker:2007a}, the observational evidence for inside-out disk growth \\citep{Munoz-Mateos:2007}, and detections of disk breaks in the distant universe \\citep{Perez:2004, Trujillo:2005}, suggest that the outer disks provide a direct view of disk assembly. Several theories for the formation of breaks have been investigated. \\Citet{van-der-Kruit:1987fk} proposed that angular momentum conservation in a collapsing, uniformly rotating cloud naturally gives rise to disk breaks at roughly 4.5 scale radii. \\Citet{van-den-Bosch:2001aa} suggested that breaks are due to angular momentum cutoffs of the cooled gas. More commonly breaks have been attributed to a threshold for star formation (SF), whether due to low gas density which stabilizes the disk \\citep{Kennicutt:1989bs}, or to a lack of a cool equilibrium ISM phase \\citep{Elmegreen:1994eu, Schaye:2004kb}. Using a semi-analytic model, \\citet{Elmegreen:2006oq} demonstrate that a double-exponential profile may result from a multi-component star formation prescription. The existence of extended UV disks \\citep[e.g][]{Thilker:2007a} and the lack of a clear correlation of H$\\alpha$ cut-offs and optical disk breaks \\citep{Pohlen:2004, Hunter:2006aa} further complicates the picture. Regardless, while a sharp SF cutoff may explain the disk truncation, it does not provide a compelling explanation for extended outer exponential components. Alternatively, \\citet{Debattista:2006wd} demonstrated that the redistribution of angular momentum by spirals during bar formation also produces realistic breaks in collisionless $N$-body simulations. In this Letter we present the first results from a series of high-resolution Smooth Particle Hydrodynamics (SPH) simulations of isolated galaxy formation aimed at exploring the formation and evolution of disk breaks and outskirts in a massive, high surface brightness galaxy without a strong central bar. Resulting breaks are analogous to downward-bending breaks seen in observations. The clear advantage of our approach over past attempts is that we use a fully self-consistent physical model of the system, making no a priori assumptions about the distribution of material in the disk. Rather, we allow the disk to grow spontaneously under the effects of gravity and gas hydrodynamics, itself influenced by star formation and feedback. The $N$-body approach (at sufficiently high mass and force resolution) ensures that we capture the dynamical processes contributing to disk evolution. Furthermore, the inclusion of prescriptions for SF and feedback allows us to make observational predictions across the break region. We show that (1) the break forms rapidly ($\\lesssim$ 1Gyr) and persists throughout the evolution of the system, moving outward as the disk mass grows; (2) the break is seeded by a sharp decrease in star formation which is caused in our simulation by a rapid decrease in the surface density of cool gas; (3) the outer disk is populated by stars that have migrated, on nearly circular orbits, from the inner disk, and consequently the break is associated with a sharp change in the radial mean stellar age profile; (4) break parameters agree with current observations. ", "conclusions": "\\label{sec:conclusions} We have shown that in a self-consistent model, where the stellar disk forms through gas cooling and subsequent star formation within a dark matter halo, breaks in the stellar surface density form through the combination of different effects. A rapid drop in the SFR seeds the break and secular evolution populates the outer exponential. In our model the SFR drop is due wholly to a drop in the surface density of gas. However a break in SFR induced by other means ({\\it e.g.} a volume density threshold or perhaps warps) would lead to similar behavior of the outer disk and stellar density break parameters. Our model properties satisfy current observational constraints, both in the statistical sense of break properties from galaxies in SDSS, as well as the much more detailed observations of breaks in stellar populations of NGC~4244. Though our model does not account for the effects of evolution in a full cosmological context, its simplicity assures that this is the minimal degree of evolution (with no interactions or formation of a significant bar) and should therefore be rather generic provided that a disk is massive enough for strong transient spirals to form. The model predicts that there should be an abrupt change in the radial mean stellar age profile coincident with the break, which can be tested with future observations." }, "0710/0710.2135_arXiv.txt": { "abstract": "IC 4406 is a large (about 100$''$ $\\times$ 30$''$) southern bipolar planetary nebula, composed of two elongated lobes extending from a bright central region, where there is evidence for the presence of a large torus of gas and dust. We show new observations of this source performed with IRAC (Spitzer Space Telescope) and the Australia Telescope Compact Array. The radio maps show that the flux from the ionized gas is concentrated in the bright central region and originates in a clumpy structure previously observed in H$\\alpha$, while in the infrared images filaments and clumps can be seen in the extended nebular envelope, the central region showing toroidal emission. Modeling of the infrared emission leads to the conclusion that several dust components are present in the nebula. ", "introduction": "\\label{introduction} IC 4406 is a well-studied southern planetary nebula. It has been imaged with several telescopes at different wavelength ranges. Near-IR images show two H$_2$ lobes \\citep{storey}, orthogonal to the nebula's major axis and $\\sim$25$''$ away from each other. These peaks are approximately coincident with the two blobs observed in H$\\alpha$+[\\ion{N}{2}] and [\\ion{O}{3}] \\citep{sahai}, interpreted as indicative of the presence of a dense equatorial torus of dust. The optical images show a central ionized region about 32$''$ in diameter. CO maps show the presence of a collimated high velocity outflow in the polar direction and with [CO]/[H$_2$]$\\approx 5\\times 10^{-6}$ and a total molecular mass in the range 0.16--3.2 M$_\\odot$ \\citep{sahai}. Hubble Space Telescope (HST) WFPC2 images in [\\ion{N}{2}], H$\\alpha$ and [\\ion{O}{3}] have revealed the existence of an intricate system of dark lane features, which led to the name of \\lq\\lq Retina Nebula\\rq\\rq~for this object \\citep{odell}. The nebula appears to be chemically homogeneous, as \\citet{corradi} found no evidence of radial variation for He, O, N, Ne, and Ar. \\citet{cox} have detected several C-rich features at mm wavelengths, such as CN, HCO$^+$, HCN and HNC, which indicate the nebula is C-rich, although a C/O ratio of 0.6 is reported by \\citet{cohen}. IC 4406 is a relatively low electron density nebula. Values in the 400-2000 cm$^{-3}$ range have been estimated using several different optical and infrared lines, with values derived by [\\ion{S}{2}] and [\\ion{O}{3}] doublets matching around 540 cm$^{-3}$ \\citep{liu,wang}. Its central star has a \\ion{He}{2} Zanstra temperature of 96800 K \\citep{phillips} and its distance is probably around 1.6 kpc \\citep{sahai}, although some authors claim it may be overestimated \\citep{odell}. \\citet{gruenwald} have modeled IC 4406 with a 3-D photoionization code and fit many observed line intensities assuming there is a torus around the central star. They find as a best fit a central star temperature of $8\\times10^4$ K, luminosity of 400 L$_\\odot$, torus density 1500 cm$^{-3}$ and nebular density 100 cm$^{-3}$. In general, comparisons of IR images of planetary nebulae, which trace the molecular gas and warm dust emission, to optical line images, which trace the ionized gas, have shown the presence of similar structures \\citep{latter}, leading to the conclusion that molecular and ionized gas spatially coexist in planetary nebulae, as well as dust grains, despite the different physical conditions these components are presumed to survive in. We have observed IC 4406 in the radio range to inspect the distribution of the ionized gas in its envelope and in the infrared to check for emission from the equatorial dust and molecular gas. In \\S\\ref{observations} we explain how we performed our observations and reduced the data; in \\S\\ref{results} we show our results and in particular in \\S\\ref{sed} how we modeled the emission in the radio and infrared ranges; \\S\\ref{nebular} compares our model results to the nebular parameter values obtained directly from the observational data; in \\S\\ref{summary} we summarize the present work. ", "conclusions": "\\label{summary} We have observed IC 4406 in the cm and 3--10 micron ranges. Our radio observations have confirmed the presence of the complicated maze of lanes already observed in H$\\alpha$ in the central region of the nebula and have enabled us to calculate several nebular parameters, whose values match the classification for this target as an evolved planetary nebula, in particular its low dust to gas mass ratio and density. IRAC imaging has revealed the presence of filaments in the nebula that were not detected in previous observations. Our IRAC measurements, combined with literature data at longer and shorter wavelengths, have enabled us to study the SED of the PN IC 4406 and reproduce it with DUSTY. This has revealed that three different dust components are needed to model the data, with temperatures ranging from 57 to 700 K. It has also been necessary to include in the model slightly larger grains than in the standard MRN composition (up to 6.5 $\\mu$m) to account for the calculated 60 $\\mu$m optical depth. The main limits of our modeled curve are the spherical geometry assumed in DUSTY and the lack of data in the mm and sub-mm ranges, which would give a constraint on the slope of the curve. As we have observed during our trials with DUSTY, the slope of the SED in the sub-mm range changes with the maximum size of the grains included in the model. Unfortunately, in this range observations are available only for a few stars so far: none for our target. We can speculate that in such a diversified dust environment, as we find in IC 4406, further lower temperature components may exist and future high sensitivity, high angular resolution observations will give a fundamental contribution to understand the physics of circumstellar envelopes in planetary nebulae." }, "0710/0710.2904_arXiv.txt": { "abstract": "Compared to planets around Sun--like stars, relatively little is known about the occurrence rate and orbital properties of planets around stars more massive than 1.3~\\msun. The apparent deficit of planets around massive stars is due to a strong selection bias against early--type dwarfs in Doppler--based planet searches. One method to circumvent the difficulties inherent to massive main--sequence stars is to instead observe them after they have evolved onto the subgiant branch. We show how the cooler atmospheres and slower rotation velocities of subgiants make them ideal proxies for F-- and A--type stars. We present the early results from our planet search that reveal a paucity of planets orbiting within 1~AU of stars more massive than 1.5~\\msun, and evidence of a rising trend in giant planet occurrence with stellar mass. ", "introduction": "A planet--bearing star can be thought of as a very bright, extremely dense remnant of a protoplanetary disk. After all, a star inherits its defining characteristics---its mass and chemical composition---from the same disk material that forms its planets. The physical characteristics of planet host stars therefore provide a crucial link between the planets we detect today and the circumstellar environments from which they formed long ago. Studying the relationships between the observed occurrence rate and orbital properties of planets as a function of stellar characteristics informs theories of planet formation, and also helps guide the target selection of future planet searches. A wealth of recent work has demonstrated that planet occurrence is strongly correlated with chemical composition \\citep{gonzalez97, santos04}; metal--rich stars are 3 times more likely to host planetary companions compared to stars with solar abundances \\citep{fischer05b}. This finding can be understood in the context of the core accretion model. Increasing the metallicity of a star/disk system increases the surface density of solid material at the disk midplane, which in turn leads to an enhanced growth rate for protoplanetary cores \\citep{ida04b, kornet05}. \\begin{figure}[ht!] \\plotone{mass_hist.eps} \\caption {\\footnotesize{Distribution of stellar masses for the target stars of the California and Carnegie Planet Search. The majority of the stars have masses between 0.7~\\msun\\ and 1.3~\\msun. \\label{mass_hist}}} \\end{figure} Another factor that enhances the surface density of material in the disk midplane is its total mass. If the mass of circumstellar disks scales with the mass of the central star, then there should be an observed correlation between planet occurrence and stellar mass \\citep{laughlin04, ida05b, kennedy07}. In principle, testing this hypothesis is fairly simple: one need only measure the fraction of stars with planets over a wide range of stellar masses. However, in practice such a study is not so straight forward given the limited range of stellar masses encompassed by most planet searches. The difficulty can be seen in Figure~\\ref{mass_hist}, which shows the distribution of stellar masses for the target stars in California and Carnegie Planet Search \\citep[CCPS; ][]{valenti05}, which is representative of most Doppler--based planet searches. Most of the stars in Figure~\\ref{mass_hist} have masses between 0.7~\\msun\\ and 1.3~\\msun. In a decidedly non--Copernican twist of nature, it turns out that stars like our Sun are ideal planet search targets. Solar--mass G and K dwarfs are slow rotators, have stable atmospheres, and are relatively bright. The fall--off toward lower stellar masses is simply because late K-- and M--type dwarfs are faint, making the acquisition of high--resolution spectra difficult without large telescope apertures \\citep{butler06b, bonfils05b, endl03}. The sharp drop at higher stellar masses is due to a separate observational bias. Stars with spectral types earlier than F8 tend to have rotationally broadened spectral features ($V\\sin{i} > 50$~\\ks; Galland et al. 2005), have fewer spectral lines due to high surface temperatures, and display a large amount of atmospheric ``jitter.'' Stellar jitter is excess velocity scatter due to surface inhomogeneities and pulsation, which can approach 50--100~\\ms\\ for mid--F stars \\citep{saar98, wright05}. These features conspire to limit the attainable radial velocity precision for early--type dwarfs to $> 50$~\\ms, rendering exceedingly difficult the detection of all but the most massive and short--period planets. One method to circumvent the observational limitations inherent to high--mass dwarfs is to observe these stars after they have evolved away from the main--sequence. After stars have expended their core hydrogen fuel sources their radii expand, and their atmospheres cool leading to an increase in the number of metal lines in the star's spectrum. Stars crossing the subgiant branch also shed a large amount of angular momentum through the coupling of stellar winds to rotationally generated magnetic fields \\citep{gray85, schrijver93, donascimento00}. The cooler atmospheres and slower rotational velocities of evolved stars lead to an increased number of narrow absorption lines in their spectra, making them much better suited for precision Doppler surveys. ", "conclusions": "" }, "0710/0710.4133_arXiv.txt": { "abstract": "If light scalar fields are present at the end of inflation, their non-equilibrium dynamics such as parametric resonance or a phase transition can produce non-Gaussian density perturbations. We show how these perturbations can be calculated using non-linear lattice field theory simulations and the separate universe approximation. In the massless preheating model, we find that some parameter values are excluded while others lead to acceptable but observable levels of non-Gaussianity. This shows that preheating can be an important factor in assessing the viability of inflationary models. ", "introduction": " ", "conclusions": "" }, "0710/0710.4075_arXiv.txt": { "abstract": "% Motivated by the emergence of multicore architectures, and the reality that parallelism is rarely used for analysis in observational astronomy, we demonstrate how general users may employ tightly-coupled multiprocessors in scriptable research calculations while requiring no special knowledge of parallel programming. Our method rests on the observation that much of the appeal of high-level vectorized languages like IDL or MatLab stems from relatively simple internal loops over regular array structures, and that these loops are highly amenable to automatic parallelization with OpenMP. We discuss how ISIS, an open-source astrophysical analysis system embedding the \\slang\\ numerical language, was easily adapted to exploit this pattern. Drawing from a common astrophysical problem, model fitting, we present beneficial speedups for several machine and compiler configurations. These results complement our previous efforts with PVM, and together lead us to believe that ISIS is the only general purpose spectroscopy system in which such a range of parallelism -- from single processors on multiple machines to multiple processors on single machines -- has been demonstrated. ", "introduction": "As noted in Noble et al (2006), parallel computation is barely used in astronomical analysis. For example, models in XSPEC (Arnaud 1996), the de facto standard X-ray spectral analysis tool, still run serially on my dual-CPU desktop. In this situation scientists tend to either turn away from models which are expensive to compute or just accept that they will run slowly. Analysis systems which do not embrace parallelism can process at most the workload of only 1 CPU, resulting in a dramatic $1/n$ underutilization of resources as more CPU cores are added. At the same time, however, astronomers are well versed in scripting, particularly with very high-level, array-oriented numerical packages like IDL, PDL, and \\slang, to name a few. They combine easy manipulation of mathematical structures of arbitrary dimension with most of the performance of compiled code, with the latter due largely to moving array traversals from the interpreted layer into lower-level code like this C fragment {\\small \\begin{verbatim} case SLANG_TIMES: for (n = 0; n $<$ na; n++) c[n] = a[n] * b[n]; \\end{verbatim} } \\noindent which provides vectorized multiplication in \\slang. This suggests that much of the strength and appeal of numerical scripting stems from relatively simple loops over regular structures. Having such loops in lower-level compiled codes also makes them ripe for parallelization with \\omp\\ on shared memory multiprocessors. Proponents contend that conceptual simplicity makes \\omp\\ more approachable than other parallel programming models, e.g. message-passing in MPI or PVM, and emphasize the added benefit of allowing single bodies of code to be used for both serial and parallel execution. For instance, preceding the above loop with \\verb+#pragma omp parallel for+ parallelizes the \\slang\\ multiplication operator; the pragma is simply ignored by a non-conforming compiler, resulting in a sequential program. ", "conclusions": "We are witnessing the arrival of serious multiprocessing capability on the desktop: multicore chip designs are making it possible for general users to access many processors. At the granularity of the operating system it will be relatively easy to make use of these extra cores, say by assigning whole programs to separate CPUs. As noted with increasing frequency of late, though, it is not as straightforward to exploit this concurrency {\\em within} individual desktop applications. In this paper we demonstrated how we have helped our research colleagues prepare for this eventuality. We have enhanced the vectorization capabilities of \\slirp, a module generator for the \\slang\\ numerical scripting language, so that wrappers may be annotated for automatic parallelization with \\omp. This lets \\slang\\ intrinsic functions be replaced with parallelized versions, at runtime, without modifying a single line of internal \\slang\\ source. We have shown how \\slang\\ operators may also be parallelized with relative ease, by identifying key loops within the interpreter source, tagging them with \\omp\\ directives and recompiling. These simple adaptations, which did not require any changes to the \\isis\\ architecture or codebase, have yielded beneficial speedups for computations actively used in astrophysical research, and allow the same numerical scripts to be used for both serial and parallel execution -- minimizing two traditional barriers to the use of parallelism by non-specialists: learning how to program for concurrency and recasting sequential algorithms in parallel form. By transparently using \\omp\\ to effect greater multiprocessor utilization we gain the freedom to explore on the desktop more challenging problems that other researchers might avoid for their prohibitive cost of computation. The \\omp\\ support now available in GCC makes the techniques espoused here a viable option for many open source numerical packages, opening the door to wider adoption of parallel computing by general practitioners. \\begin{figure*}[t] \\centering \\includegraphics[angle=-90,scale=0.22]{P10.5_3.eps} \\hspace*{6mm} \\includegraphics[angle=-90,scale=0.22]{P10.5_4.eps} \\caption{Aggregate speedup of the Weibull fit function due to the parallelized operators and functions discussed above. Left: \\lint, with inflection point at 1907 elements. Right: \\solf, with inflection point at 384 elements. } \\label{P10.5-fig-2} \\end{figure*}" }, "0710/0710.1650_arXiv.txt": { "abstract": "We report on an investigation of the SBS 1520+530 gravitational lens system. We have used archival HST imaging, Keck spectroscopic data, and Keck adaptive-optics imaging to study the lensing galaxy and its environment. The AO imaging has allowed us to fix the lens galaxy properties with a high degree of accuracy when performing the lens modeling, and the data indicate that the lens has an elliptical morphology and perhaps a disk. The new spectroscopic data suggest that previous determinations of the lens redshift may be incorrect, and we report an updated, though inconclusive, value $z_{lens} = 0.761$. We have also spectroscopically confirmed the existence of several galaxy groups at approximately the redshift of the lens system. We create new models of the lens system that explicitly account for the environment of the lens, and we also include improved constraints on the lensing galaxy from our adaptive-optics imaging. Lens models created with these new data can be well-fit with a steeper than isothermal mass slope ($\\alpha = 2.29$, where $\\rho \\propto r^{-\\alpha}$) if $H_0$ is fixed at 72\\hunit; isothermal models require $H_0 \\sim 50$\\hunit. The steepened profile may indicate that the lens is in a transient perturbed state caused by interactions with a nearby galaxy. ", "introduction": "The strong gravitational lens system SBS 1520+530 (hereafter SBS1520) was first investigated by \\citet{chavushyan}, who found that the system consists of a pair of images of a broad absorption line quasar ($z_{src} = 1.855$) separated by 1\\farcs6. The lensing galaxy was soon detected with adaptive optics (AO) imaging \\citep{crampton} and was assumed to be at the redshift of one of two absorption line systems seen in the spectra of the quasar images. \\citet{burud} attempted to deconvolve the lens spectrum from the quasar spectra and found the redshift of the lens to be $z_{lens} = 0.717$. This redshift is broadly consistent with a photometric redshift determined by \\citet{faure}. Furthermore, the lens was found to lie along the line of sight to a photometrically identified cluster of galaxies that is expected to be at approximately the same redshift as the lens \\citep{burud,faure}. Optical monitoring campaigns have led to the measurement of a time delay between the quasar images of $\\sim 130$ days \\citep{burud,khamitov}. This time delay provides an additional constraint for determining the mass slope of the lens galaxy \\citep[e.g.,][]{rusin} and allows a value of the Hubble Constant to be determined for the system \\citep[$H_0 = 51$\\hunit assuming an isothermal mass profile;][]{burud}. Note, however, that the mass slope and, thus, the value of $H_0$ depend on the environment surrounding the lens system \\citep[e.g.,][]{dobke,auger}. An incorrect understanding of the mass distribution and environment of the lens might account for the anomalously low value of $H_0$ obtained for SBS1520 compared to other lens systems \\citep[e.g.,][]{koopmans03,york} and the WMAP \\citep[][]{spergel} and \\em{Hubble Space Telescope} (\\em{HST}) Key Project \\citep[][]{freedman} results. In this paper we present new Keck AO and archival \\emph{HST} imaging of the lens system that indicates the lensing galaxy may have a disk component. We also present a spectroscopic investigation of the lens environment and provide a new analysis of the lensing galaxy's spectrum which results in an updated lens redshift of $z_{lens} = 0.761$. We discuss the implications of these new observational data on previous analyses performed with SBS1520 and suggest that, in spite of some complexity, this lens provides an interesting platform to investigate dark matter interactions in dense environments. ", "conclusions": "We have obtained deep AO imaging and optical spectroscopy of the time-delay lens SBS1520. The AO imaging has allowed us to fix the lens galaxy properties with a high degree of precision when performing the lens modeling, and the data indicate that the lens has an elliptical morphology and perhaps a disk. The new spectroscopic data suggest that previous determinations of the lens redshift may be incorrect, and the data also allow us to quantify the lensing contribution of several groups in the immediate foreground and background of the lens. Lens models created with these new data can be well-fit with a steeper than isothermal mass slope ($\\alpha = 2.29$) if $H_0$ is fixed at 72\\hunit; isothermal models require $H_0 \\sim 50$\\hunit. \\citet{dobke} found that galaxies in overdense environments might have steeper than isothermal mass slopes caused by interactions with other galaxies \\citep[also see][]{augerslacs}. This suggests an interpretation that we are observing transient steepening of the mass profile due to galaxy-galaxy interactions and indicates that other lens systems that have obtained anomalously low values of $H_0$ may lie in overdense regions and near an interacting galaxy \\citep[e.g., B1600+434;][]{koopmans1600,auger}. Alternatively, SBS1520 can be modeled in a manner consistent with an isothermal profile and $H_0 = 64$\\hunit~if if the lens is modeled by a pixelated mass distribution and jointly modeled with other lens systems \\citep{read}. These models indicate that twisting ellipticity, triaxial projection effects, or other shape degeneracies may be effecting the parametric analyses of SBS1520 \\citep{saha}. However, there are still several ambiguities in the data that need to be resolved before making definitive claims about the profile of SBS1520, particularly in the context of the interaction-induced steepening scenario. While we have argued that the lens redshift is likely to be $z = 0.761$ and not $z = 0.717$, the data are not conclusive. Furthermore, our modeling has assumed that all of the neighbor galaxies are at the group redshift; if this is the case, the $z = 0.76$ group centroid would be pulled closer to the lens and the group would therefore provide a larger contribution to the lens model. If this is not the case, the neighboring galaxies might have a smaller impact on the lens model. This is particularly important for Galaxy M, as this is the neighboring galaxy that most affects the lens model but also has colors least like the lens galaxy compared to the other field galaxies. It is also important to verify that at least one of the neighboring galaxies is at the same redshift as the lens because this is a requirement of the interaction-driven steepening hypothesis. Finally, obtaining a dynamical estimate of the lens mass would help to further constrain models and potentially distinguish between shape degeneracies and the mass slope degeneracy." }, "0710/0710.5073_arXiv.txt": { "abstract": "Due to the non-commutation of spatial averaging and temporal evolution, inhomogeneities and anisotropies (cosmic structures) influence the evolution of the averaged Universe via the cosmological backreaction mechanism. We study the backreaction effect as a function of averaging scale in a perturbative approach up to higher orders. We calculate the hierarchy of the critical scales, at which $10\\%$ effects show up from averaging at different orders. The dominant contribution comes from the averaged spatial curvature, observable up to scales of $\\sim 200~\\mbox{Mpc}$. The cosmic variance of the local Hubble rate is $10\\%$ $(5\\%)$ for spherical regions of radius 40 $(60)~\\mbox{Mpc}$. We compare our result to the one from Newtonian cosmology and Hubble Space Telescope Key Project data. ", "introduction": " ", "conclusions": "" }, "0710/0710.5303_arXiv.txt": { "abstract": "We have updated predictions for high energy neutrino and antineutrino charged current cross-sections within the conventional DGLAP formalism of NLO QCD using a modern PDF fit to HERA data, which also accounts in a systematic way for PDF uncertainties deriving from both model uncertainties and from the experimental uncertainties of the input data sets. Furthermore the PDFs are determined using an improved treatment of heavy quark thresholds. A measurement of the neutrino cross-section much below these predictions would signal the need for extension of the conventional formalism as in BFKL resummation, or even gluon recombination effects as in the colour glass condensate model. ", "introduction": "\\label{sec:intro} Predictions of neutrino cross-sections at high energies have sizeable uncertainties which derive largely from the measurement uncertainties on the parton distribution functions (PDFs) of the nucleon. In the framework of the quark-parton model, high energy scattering accesses very large values of $Q^2$, the invariant mass of the exchanged vector boson, and very small values of Bjorken $x$, the fraction of the momentum of the incoming nucleon taken by the struck quark. Thus when evaluating uncertainties on high energy neutrino cross-sections it is important to use the most up to date information from the experiments at HERA, which have accessed the lowest-$x$ and highest $Q^2$ scales to date. The present paper uses the formalism of the ZEUS-S global PDF fits~\\cite{Chekanov:2002pv}, updated to include {\\em all} the HERA-I data. Conventional PDF fits use the Next-to-leading-order (NLO) Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) formalism~\\cite{Altarelli:1977zs,Gribov:1972ri,Lipatov:1974qm,Dokshitzer:1977sg} of QCD to make predictions for deep inelastic scattering (DIS) cross-sections of leptons on hadrons. At low-$x$ where the gluon density is rising rapidly it is probably necessary to go beyond the DGLAP formalism in order to sum $\\ln(1/x)$ diagrams, as in the Balitsky-Fadin-Kuraev-Lipatov (BFKL) formalism~\\cite{Kuraev:1977fs,Balitsky:1978ic,Lipatov:1985uk} (for recent work see~\\cite{White:2006xv,Altarelli:2005ni,Ciafaloni:2006yk}). An alternative approach is to consider non-linear terms which describe gluon recombination as in the colour glass condensate model~\\cite{Iancu:2003xm} which has had considerable success in explaining RHIC data~\\cite{JalilianMarian:2005jf}. A recent suggestion is to use a structure function consistent with HERA data that saturates the Froissart unitarity bound and thus predicts a $\\ln^2 s$ dependence of the cross-section~\\cite{Berger:2007ic}. Such approaches are beyond the scope of the present paper, which is concerned with the more modest goal of estimating the uncertainties on high energy neutrino cross-sections which are compatible with the conventional NLO DGLAP formalism. The motivation is to provide an update on the neutrino cross-sections in the literature \\cite{Gandhi:1998ri} which are widely used e.g. for estimating event rates in neutrino telescopes such as Baikal \\cite{Antipin:2007zz}, ANTARES \\cite{Aslanides:1999vq} and IceCube \\cite{icecube}, cosmic ray observatories such as HiRes \\cite{Martens:2007ff} and Auger \\cite{auger}, and radio detectors such as GLUE \\cite{Gorham:2003da}, FORTE \\cite{Lehtinen:2003xv}, RICE \\cite{Kravchenko:2002mm} and ANITA \\cite{Barwick:2005hn}. As a corollary, if cross-sections much outside these limits are observed, it would be a clear signal of the need for extensions to conventional formalism. To date no unambiguous signals which require such extensions have been observed. The prospect for measuring the cross-section using very high energy cosmic neutrinos in order to distinguish between theoretical suggestions for gluon dynamics at low $x$ has been discussed by us elsewhere \\cite{Anchordoqui:2006ta}. Previous work on estimating high energy neutrino cross-sections~\\cite{Gandhi:1998ri} used PDF sets which no longer fit modern data from HERA~\\cite{Tung:2004rw} and an {\\em ad hoc} procedure for estimating PDF uncertainties. The present work improves on this in several respects. Firstly, we use a recent PDF analysis which includes data from all HERA-I running~\\cite{Chekanov:2002pv}. Secondly, we take a consistent approach to PDF uncertainties --- both model uncertainties and, more importantly, the uncertainties which derive from the correlated systematic errors of the input data sets \\cite{CooperSarkar:2002yx}. Thirdly, we use NLO rather than LO calculations throughout. Fourthly, we use a general-mass variable flavour number scheme~\\cite{Thorne:1997ga,Thorne:2006qt} to treat heavy quark thresholds. ", "conclusions": "\\label{sec:conc} We have calculated the charged current neutrino cross-section at NLO in the Standard Model using the best available DIS data along with a careful estimate of the associated uncertainties. As mentioned earlier, there are further uncertainties associated with QCD effects at very low $x$ which are not addressed in the DGLAP formalism. When $x$ is sufficiently small that $\\alpha_\\mathrm{s}\\,\\ln (1/x) \\sim 1$, it is necessary to resum these large logarithms using the BFKL formalism. Whereas such calculations at leading-log suggest an even steeper rise of the gluon structure function at low $x$ (which would imply a higher $\\nu-N$ cross-section), this rise is not so dramatic at next-to-leading-log; for a recent application of NLL BFKL resummation to deep inelastic scattering see~\\cite{White:2006yh}. Moreover both the DGLAP and the BFKL formalisms neglect non-linear screening effects due to gluon recombination which may lead to saturation of the gluon structure function. This has been modelled in the colour dipole framework in which DIS at low $x$ is viewed as the interaction of the $q \\bar q$ dipole to which the gauge bosons fluctuate. An unified BFKL/DGLAP calculation \\cite{Kwiecinski:1998yf} supplemented by estimates of screening and nuclear shadowing effects, predicts a {\\em decrease} of the $\\nu-N$ cross-section by $20-100\\%$ at very high energies $E_\\nu \\sim 10^8-10^{12}$~GeV~\\cite{Kutak:2003bd}. An alternative approach uses the colour glass condensate formalism~\\cite{Iancu:2003xm} and predicts a similar suppression when a dipole model~\\cite{Kharzeev:2004yx} which fits data from RHIC is used~\\cite{Henley:2005ms}. The predicted cross-section is even lower~\\cite{Henley:2005ms} if a different dipole model~\\cite{Bartels:2002cj} developed to fit the HERA data is used and the gluon distribution is assumed to decrease for $x < 10^{-5}$. Other possibilities for the behaviour of the high energy $\\nu-N$ cross-section have also been discussed~\\cite{Berger:2007ic,Jalilian-Marian:2003wf}. Detectors for UHE cosmic neutrinos would be able to probe such new physics if they can establish deviations from the perturbative DGLAP prediction. Hence we recommend our calculated values for estimation of the baseline event rates in neutrino telescopes and for use in event generators such as ANIS \\cite{Gazizov:2004va}. While the expected neutrino fluxes (e.g. from the sources of the observed high energy cosmic rays) are rather uncertain, experiments can in principle exploit the different dependence on the cross-section of the rate of Earth-skimming and quasi-horizontal events \\cite{Anchordoqui:2006ta}." }, "0710/0710.5629_arXiv.txt": { "abstract": "The specifications of the Atacama Large Millimeter Array (ALMA) have placed stringent requirements on the mechanical performance of its antennas. As part of the evaluation process of the VertexRSI and Alcatel EIE Consortium (AEC) ALMA prototype antennas, measurements of the path length, thermal, and azimuth bearing performance were made under a variety of weather conditions and observing modes. The results of mechanical measurements, reported here, are compared to the antenna specifications. ", "introduction": "\\PARstart{T}{he} Atacama Large Millimeter Array (ALMA) for astronomical observations at millimeter and sub--millimeter wavelengths (up to the Terra--Hertz region) needs antennas of high mechanical precision and of understandable and predictable behaviour. This behaviour must be established for structural deformations due to gravity, temperature changes, and wind loads. This means, in particular, that a high reflector surface precision, pointing and phase stability must be maintained under all motions of tracking and mapping. We present a summary of tests of the mechanical and thermal behaviour of the 12m diameter VertexRSI and AEC ALMA prototype antennas, built at the VLA site (2000m altitude), New Mexico, USA. The tests were made at several intervals between March 2003 and April 2004, and concentrated primarily on the verification of the antenna specifications, of path length variations and parameters which influence the pointing. The data were also analyzed to understand the general behaviour of the antennas. In the investigation we have paid attention to the fact that variations in the behaviour of the antennas may be predictable or sporadic. We believe that repeatable and/or predictable variations can to a large extent be considered in the pointing model, or any other correction device. The antennas were tested in stationary position, under sidereal tracking, On--the--Fly (OTF) mapping, and in Fast--Switching mode (FSW). A large amount of data was collected during commissioning and thus refer to all types of tracking, OTF, FSW, and unintended 'shaking'. A more extended summary of these test results was reported by the Antenna Evaluation Group to the National Radio Astronomy Observatory (NRAO) and European Southern Observatory (ESO) (which forms the ALMA partnership) for selection of the ALMA production antenna(s). An overview of these performance results was presented in \\cite{Mangum2006}. The current paper presents a more detailed analysis of the mechanical performance measurements made of the ALMA prototype antennas. ", "conclusions": "The measurements indicate that the path length specifications are fulfilled on both antennas, at least during time intervals of 1/2 to 1 hour. The path length variation is primarily due to unavoidable residual thermal dilatation of the (insulated) antenna steel components, and may span $\\sim$\\,200\\,$\\mu$m within a day. The path length variation can be predicted with high precision from temperature measurements at a few positions of the steel components, either used in empirical relations or the finite element model. It is expected that identical antennas will experience similar temperature variations of the ambient air so that the differential effect may even be smaller than stated here. Wind at speeds below the specification limit (9\\,m/s), and OTF and FSW motions of the antennas, do not degrade the path length stability. As far as possible to measure, the antennas show similar behaviour of damping of the thermal environment, \\ie\\ the ambient air temperature and the solar radiation. The BUS of the VertexRSI antenna shows a good temperature homogeneity, even under full exposure to Sun shine. Altough the AZ bearings have a higher order azimuth dependent wobble, the effect can be considered in the pointing model with an accuracy better than 0.6 arcsecond. On the VertexRSI prototype antenna, the wobble was very stable with time." }, "0710/0710.2009_arXiv.txt": { "abstract": " ", "introduction": "\\label{sec:intro} There has been much interest in recent years in cosmological applications of string theory. The availability of precision data relating to the cosmic microwave background (CMB) opens up the possibility of constraining the phenomenology of string--inspired models, in particular regarding the dynamics of the inflationary epoch. Essentially, signatures of the extreme conditions governing the behaviour of the early universe are amplified by inflationary expansion and can be accessible by modern day observations. The typical energy scales involved far exceed those obtainable in particle collider experiments. In this paper we consider a particular inflation scenario, the Dirac--Born--Infeld (DBI) model of~\\cite{Silverstein:2003hf,Alishahiha:2004eh}. In the usual formulation of this setup, a D3 brane rolls down a warped throat towards a $\\overline{\\text{D3}}$ brane where the interbrane separation is identified with the inflaton field. The brane moves relativistically, but its speed is curtailed by the warped geometry so that the potential energy dominates and inflation can occur. The observable consequences of this model depend of course upon the choice of solution for the warped throat. The original DBI proposal used a AdS--like geometry, with an artificially imposed cutoff where the throat joins the bulk geometry, which in principle is unknown although irrelevant for cosmological implications. Such a geometry is unstable, and more properly one should consider a solution of the supergravity field equations with non--trivial D--brane fluxes to stabilise the compact geometry by dynamically generating a cut--off. Not many such solutions are known analytically, but two examples are the Klebanov--Strassler~\\cite{Klebanov:2000hb} (KS) and Maldacena--Nu{\\~n}ez~\\cite{Maldacena:2000yy} (MN) results.\\\\ Several investigations of DBI inflation from the KS geometry abound in the literature~\\cite{Kecskemeti:2006cg,Shiu:2006kj,Easson:2007dh,Shandera:2006ax,Bean:2007hc}, including systematic comparisons to current cosmological data. In this paper we show that it is also possible to examine the inflationary consequences of the BGMPZ geometries introduced in~\\cite{Butti:2004pk}. These are a one--parameter family of geometries describing a deformation of the KS solution\\footnote{The BGMPZ solutions originally arose in a different context, that of the AdS / CFT correspondence, as the baryonic branch of the KS solution.}, and interpolating smoothly between the KS and MN backgrounds. Although the metric is not known fully analytically, one can solve for it numerically and thus obtain results for numerous observables. The hope is that these can be compared with experimental data for quantities such as spectral indices, which may then allow one to extract geometrical information about the throat--geometry. To this aim, we show examples of how these quantities vary as a function of the BGMPZ parameter $\\xi$ which characterises the family of solutions, at typical points in parameter space. We note that inflationary consequences of these geometries were also considered in~\\cite{Dymarsky:2005xt}, for slow-roll brane inflation rather than the DBI setup considered in this paper.\\\\ The structure of the paper is as follows. In section \\ref{sec:dbi} we summarise the salient details of DBI inflation relevant for our purposes, followed by a brief explanation of BGMPZ backgrounds. In section \\ref{sec:calc} we explain the details of the numerical procedure adopted in solving the BGMPZ equations. Example results are presented in section \\ref{sec:results}, and in section \\ref{sec:discuss} we discuss our results before concluding. ", "conclusions": "\\label{sec:discuss} In this paper we have explicitly demonstrated the possibility of calculating cosmological observables in the DBI inflation setup using a one parameter family of type IIB supergravity solutions that describe the geometry of a warped throat, the BGMPZ solutions of Butti et al.~\\cite{Butti:2004pk} that interpolate smoothly between the Klebanov--Strassler and the Maldacena--Nu{\\~n}ez solution. We have provided examples of cosmological parameters, namely spectral indices, that can be calculated from the underlying geometry.\\\\ The solution for the metric of the geometries in question is not possible analytically. Therefore numerical methods have been used and shown to provide an adequate representation of the metric in terms of numerical precision. Instabilities in the derivatives entering the equations have been dealt with by matching to known power series expansions of the metric functions at asymptotically small and large values of the radial coordinate.\\\\ We presented warp factors for two almost extremal ends of the family of solutions parametrised by~$\\xi$: the Klebanov--Strassler throat ($\\xi=1/2$), and a geometry close to the Maldacena--Nu{\\~ n}ez solution ($\\xi=0.167$). The qualitative behaviour of the warp factors is seen to be different. Both solutions show a flat warp factor at low values of the rescaled radial coordinate $\\phi$, corresponding to a dynamically generated cutoff, and an asymptotically AdS behaviour at large $\\phi$ values as expected. However, the warp factor as it moves away in $\\xi$ from the KS solution develops a shoulder at intermediate values, whose slope is in general different from the eventual asymptotic behaviour. This can be effectively parameterised by a function of form (\\ref{fpar2}), thus explicitly demonstrating that phenomenological fits of DBI inflation to cosmological data that assume a warp factor of this form indeed correspond to exactly realisable warped throat geometries in a known string theory compactification.\\\\ We presented examples of scalar spectral indices and the ratio of tensor to scalar modes calculated using the different geometries. The different qualitative behaviour observed in the warp factors carries through to the spectral indices, and a quantitative difference between the values of the indices from different solutions at 55 $e$--folds before the end of inflation is potentially measurable. Constraints on non-Gaussianity and measured values of the scalar spectral index could already be used to rule out some regions of $\\xi$, when it is considered alongside the rest of the DBI parameter space. Very generically, we have found that the amount of non--Gaussianity increases as one moves away from the KS solution (all other parameters being equal).\\\\ Since the warping changes qualitatively as one varies $\\xi$, it is also conceivable that that certain constraints on the parameter space may be relaxed by including $\\xi$ as an additional parameter with respect to existing analyses such as~\\cite{Lorenz:2007ze}. This issue certainly deserves further study. \\subsection*{Acknowledgements} We would like to thank Gary Shiu for initiating this project and for very helpful correspondence. We acknowledge informative discussions with Marieke Postma. Our research is supported by the Dutch Foundation for Fundamental Research of Matter (FOM)." }, "0710/0710.3112_arXiv.txt": { "abstract": "We study CP-violation effects when neutrinos are present in dense matter, such as outside the proto-neutron star formed in a core-collapse supernova. Using general arguments based on the Standard Model, we confirm that there are no CP-violating effects at the tree level on the electron neutrino and anti-neutrino fluxes in a core-collapse supernova. On the other hand significant effects can be obtained for muon and tau neutrinos even at the tree level. We show that CP violating effects can be present in the supernova electron (anti)neutrino fluxes as well, if muon and tau neutrinos have different fluxes at the neutrinosphere. Such differences could arise due to physics beyond the Standard Model, such as the presence of flavor-changing interactions. ", "introduction": "Recent results from solar, atmospheric and reactor experiments have significantly improved our knowledge of the neutrino mass differences and of two of the mixing angles. If the remaining mixing angle, $\\theta_{13}$, is relatively large there is a possibility that violation of CP symmetry may be observable in the neutrino sector. Currently planned and future experiments will have improved sensitivities to the value of this angle (see e.g. \\cite{Guo:2007ug,Ardellier:2006mn,Kato:2007zz}). Effects of CP-violation in accelerator neutrino oscillation experiments have been extensively investigated \\cite{Dick:1999ed,Peres:2003wd,Ishitsuka:2005qi,Barger:1980jm,Karagiorgi:2006jf,Volpe:2006in}. The discovery of a non zero Dirac delta phase might help our understanding of the observed matter-antimatter asymmetry of the universe \\cite{Pascoli:2006ie,Pascoli:2006ci,Mohapatra:2006se,Barger:2003gt}. Besides studies on terrestrial experiments with man-made sources, a few recent works have addressed CP-violation with neutrinos from astrophysical sources (see e.g. \\cite{Akhmedov:2002zj,Winter:2006ce}). The purpose of the present paper is to explore possible effects coming from the CP-violating phase in dense matter, such as that encountered in core-collapse supernovae. Core-collapse supernovae occur following the stages of nuclear burning during stellar evolution after an iron core is formed. The iron cores formed during the evolution of massive stars are supported by the electron degeneracy pressure and hence are unstable against a collapse during which most of the matter is neutronized. Once the density exceeds the nuclear density this collapse is halted. Rebounding pressure waves break out into a shock wave near the sonic point where the density reaches the nuclear density. Evolution of this shock wave and whether it produces an explosion is a point of current investigations. However, it is observed that the newly-formed hot proto-neutron star cools by neutrino emission. Essentially the entire gravitational binding energy of eight or more solar mass star is radiated away in neutrinos. Although the initial collapse is a very orderly (i.e. low entropy) process, during the cooling stage at later times a neutrino-driven wind heats the neutron-rich material to high entropies \\cite{Woosley:1992ek,Woosley:1994ux,Takahashi:1994yz}. Neutrino interactions play a very important role in the evolution of core-collapse supernovae and in determining the element abundance \\cite{Balantekin:2003ip}. Neutrino heating is a possible mechanism for reheating the stalled shock \\cite{Bethe:1984ux}. A good fraction of the heavier nuclei were formed in the rapid neutron capture (r-process) nucleosynthesis scenario \\cite{Burbidge:1957vc}. Core-collapse supernovae are one of the possible sites for the r-process nucleosynthesis. A key quantity for determining the r-process yields is the neutron-to-seed nucleus ratio, determined by the neutron-to-proton ratio, which is controlled by the neutrino fluxes. In addition, recent work indicates that neutrino-neutrino interactions plays a potentially very significant role in supernovae \\cite{Samuel:1993uw,Sigl:1992fn,Balantekin:2004ug,Duan:2006jv,Balantekin:2006tg,Hannestad:2006nj}. In this paper we study CP violation aspects in the core-collapse supernova environment. We first analyze analytically and in general terms, how the neutrino propagation equations and the evolution operator are modified in matter, in presence of a non-zero Dirac delta phase. We obtain a general formula which is valid for any matter density profile\\footnote{Such findings are in agreement with what was found in Ref. \\cite{Yokomakura:2002av}.}. In particular we demonstrate that, as in vacuum, the electron (anti)neutrino fluxes are independent of the phase $\\delta$, if mu and tau neutrinos have the same fluxes at the neutrinosphere in the supernova\\footnote{A remark on this aspect was made in \\cite{Yoshida:2006sk}.}. On the other hand the electron (anti)neutrino fluxes will depend on $\\delta$, if mu and tau neutrinos have different fluxes at the neutrinosphere, at variance with what was found in \\cite{Akhmedov:2002zj}. We present numerical calculations on possible CP violation effects on the mu and tau neutrino fluxes as well as on the electron (anti-)neutrino flux and the electron fraction. The latter can only appear if physics beyond the Standard Model, such as flavor changing interactions, induces differences on the mu and tau neutrino initial total luminosities and/or temperatures. Finally we calculate the effects from the CP violating phase on the number of events in an observatory on earth. The plan of this paper is as follows. In Section II we present the general formalism to describe the neutrino evolution in presence of the $\\delta$ phase. The formulas concerning neutrino fluxes and the electron fraction in the supernova environment are recalled in Section III. Numerical results on these quantities are presented in Section IV. Conclusions are made in Section V. ", "conclusions": "\\noindent In this work we have analyzed possible effects induced by the CP violating Dirac phase in a dense environment such as the core-collapse supernovae. Our major result are that in matter: i) significant effects are found on the muon and tau neutrino fluxes for a non-zero CP violating phase; ii) important effects are also found on the electron (anti)neutrino fluxes if the $\\nu_{\\mu}$ and $\\nu_{\\tau}$ neutrino fluxes differ at the neutrinosphere. On the other hand the usual assumption of ignoring the CP violating phase made in the literature is justified if contributions from physics beyond the Standard Model is small and the $\\nu_{\\mu}$ and $\\nu_{\\tau}$ fluxes are equal at emission. We have calculated the events in an observatory on earth and shown that effects at the level of $5 \\%$ are present on the number of events as a function of neutrino energy. Recent calculations have shown that the inclusion of neutrino-neutrino interaction introduces new features in the neutrino propagation in supernovae. A detailed study of the neutrino evolution with the CP violating phase, the neutrino-neutrino interaction as well as loop induced neutrino refractive indices will be the object of further work." }, "0710/0710.3748_arXiv.txt": { "abstract": "We used multiwavelength data (H{\\sc i}, FUV, NUV, R) to search for evidence of star formation in the intragroup medium of the Hickson Compact Group 100. We find that young star-forming regions are located in the intergalactic H{\\sc i} clouds of the compact group which extend to over 130 kpc away from the main galaxies. A tidal dwarf galaxy candidate is located in the densest region of the H{\\sc i} tail, 61 kpc from the brightest group member and its age is estimated to be only 3.3 Myr. Fifteen other intragroup H{\\sc ii} regions and TDG candidates are detected in the GALEX FUV image and within a field 10$'$$\\times$10$'$ encompassing the H{\\sc i} tail. They have ages $<$200 Myr, H{\\sc i} masses of 10$^{9.2-10.4}$ M$_{\\odot}$, 0.001$<$ SFR $<$0.01 M$_{\\odot}$ yr$^{-1}$, and stellar masses 10$^{4.3}$--10$^{6.5}$ M$_{\\odot}$. The H{\\sc i} clouds to which many of them are associated have column densities about one order of magnitude lower than N(H{\\sc i})$\\sim$10$^{21}$ cm$^{-2}$. ", "introduction": "The environment of galaxies play an important role on determining their overall properties. One of the most interesting environmental effects is seen in interacting systems which contain stripped H{\\sc i} gas in the intragroup/intergalactic medium, instead of around galaxies. Three independent studies, Mendes de Oliveira et al. (2004, for the Stephan's Quintet), Mendes de Oliveira et al. (2006, for HCG 31), Ryan-Weber et al. (2004, for NGC 1533, HCG 16, ESO 149-G003) and Oosterloo et al. (2004, for the loose group around NGC 1490) have surveyed systems with H{\\sc i} intergalactic clouds and have shown that these are associated with actively star forming regions, the so-called {\\it intergalactic H{\\sc ii} regions}. These objects seem to be similar to the H{\\sc ii} regions in our Milky Way, but are located in the intragroup medium and have high metallicities (close to solar in most cases). The fate of these types of objects and their importance in galaxy formation and evolution, enrichment of the intergalactic medium and globular cluster formation is still debatable. In order to better address these issues, we embarked on a multiwavelength analysis of the environment of interacting systems and present here the results based on FUV, NUV, optical, and H{\\sc i} data of a Hickson compact group, HCG100. The last group in Hickson's catalog of compact groups of galaxies (Hickson 1982) is at v$_{\\rm R}$ = 5336 km/s (z=0.0178, Hickson et al. 1992) and it is formed by four late-type galaxies with accordant redshifts: a bright central Sb galaxy (HCG100a), an irregular galaxy with an optical tidal tail (HCG100b), a late-type barred spiral (HCG100c) with central H$\\alpha$ emission (Vilchez \\& Iglesias-P\\'aramo 1998) and a late-type edge-on spiral (HCG100d). HCG100a, b and c show strong evidence of interaction, demonstrated not only by peculiarities in their morphologies but also in their velocity fields (Plana et al. 2003). H100b shows a tidal tail connecting with a faint galaxy not originally classified as a member of the group by Hickson. Past encounters are also confirmed by an extended H{\\sc i} tail (Verdes-Montenegro et al. 2006). Therefore, the well-advanced dynamical state of HCG100 makes it an ideal target for searching for intergalactic H{\\sc ii} regions that might have been triggered by tidal interaction. This paper is organized as follows: \\S 2 describes the data, \\S 3 discusses the age, masses and star formation rate estimates, \\S 3.1 and \\S 3.2 compare the intragroup regions in HCG100 with other intergalactic H{\\sc ii} regions and high$-z$ UV-luminous galaxies. Finally, \\S4 summarizes the main conclusions. Throughout this paper, we use a cosmology with $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda}=0.7$~and $h=0.7$. Magnitudes are given in the AB-system and the adopted distance to HCG100 is 76.3 Mpc. \\section {The data} HCG100 was observed with the Galaxy Evolution Explorer (GALEX) mission in the far and near ultraviolet (FUV $\\lambda$$_{\\rm eff}$=1528\\AA, NUV $\\lambda$$_{\\rm eff}$=2271\\AA) as part of our Cycle 1 program (GI\\#73) to observe compact groups of galaxies. In Fig. \\ref{fuvhi_10arcmin} and Fig.\\ref{nuvhi_10arcmin} we show a cutout of a 10$'$$\\times$ 10$'$ window of the FUV and NUV images (GALEX field of view is 1$^{\\circ}$.28 and 1$^{\\circ}$.24 in FUV and NUV, respectively and pixel scale is 1.5 arcsec pixel$^{-1}$) together with the distribution of atomic hydrogen H{\\sc i} (PI Verdes-Montenegro). We have also obtained an R-band image with the CTIO Blanco 4m telescope and a mosaic II CCD imager. Three 300s images of the group were taken, at a median seeing of 1.1$''$. Each frame covered a 40 arcmin field on a side, at a pixel scale of 0.27$''$/pixel, but we only used a 10 arcmin field around the object, given that this was the area of interest. The data was bias subtracted and flat fielded using standard procedures (using the package mscred in IRAF\\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.}). A flatfield was constructed from a combination of dark sky frames and twilight flats which worked well, given that the background, after flatfielding was done, showed an rms variation no greater than 1\\% over the whole field. The night was not photometric and therefore no calibration stars were observed. Instead, we chose to use two galaxies previously observed by Rubin et al. (1991), HCG 100c and HCG 100d, for obtaining the zero point of the photometry (see below). \\subsection{Flux Calibration} FUV and NUV fluxes were calculated using Morrissey et al. (2005) m$_{\\lambda}$=-2.5 log[F$_{\\lambda}$/a$_{\\lambda}$] + b$_{\\lambda}$, where a$_{FUV}$ = 1.4 $\\times$ 10$^{-15}$ erg s$^{-1}$ cm$^{-2}$ \\AA$^{-1}$, a$_{NUV}$=2.06$\\times$ 10$^{-16}$ erg s$^{-1}$ cm$^{-2}$ \\AA$^{-1}$, b$_{FUV}$=18.82 and b$_{NUV}$=20.08 for FUV and NUV, respectively. Fluxes were multiplied by the effective filter bandpass ($\\Delta$$\\lambda$$_{FUV}$=269\\AA\\, and $\\Delta$$\\lambda$$_{NUV}$=616\\AA) to give units of erg s$^{-1}$ cm$^{-2}$ and luminosities were calculated for a distance D=76.3 Mpc. The R-band image was calibrated to match Rubin et al. (1991) magnitudes for galaxies HCG100c and HCG100d. Total integrated R magnitudes of these objects are given in their Table 3 (no errors were quoted), and these were compared with the total instrumental magnitudes we obtained for these galaxies using two methods: (1) through aperture photometry in IRAF (using the task daophot.phot) and (2) using the program SExtractor (Bertin and Arnouts et al. 1996). The zero points obtained for each of these objects, and for each method, agreed within 0.1 mag, and we used an average of the values as our final zero point. The lack of a proper calibration, does not significantly change our results as shown in \\S~3. The H{\\sc i} data taken with the VLA (configuration D, PI Verdes-Montenegro) is not able to resolve the intragroup objects since the beamwidth is 61.0$''$$\\times$55.23$''$ (22.6~kpc $\\times$ 20.4~kpc at 76.3~Mpc) and objects have sizes $<$16$''$. Therefore, our measurements of the H{\\sc i} fluxes were centered on the FUV detections but measured within one beam. Fluxes were converted into mass by using the relation: \\begin{equation} M_{HI}=2.356 \\times 10^5 F_{HI} D^2 \\end{equation} where $D$ is the distance to the group in Mpc, $F_{HI}$ is the H{\\sc i} flux in Jy km/s. Column Densities were calculated by applying the following relation \\begin{equation} N_{HI}=1.82\\times 10^{18} (\\lambda ^2 / (2.65 \\Theta ^2) F \\end{equation} where $\\Theta$=beamwidth in arcmin$^{2}$, $\\lambda$ is the wavelength (21cm) and F is the H{\\sc i} flux density in Jy/beam km/s (Rohlfs \\& Wilson 2000). \\subsection{Source Detection} We used SExtractor (SE, Bertin \\& Arnouts 1996) to detect sources in the FUV and matched that catalog with the NUV and R-band catalogs. Therefore, only objects with FUV detections were included in our final catalog. In this paper we concentrate on all objects which were detected in a region of 10$'$$\\times$ 10$'$ centered on the H{\\sc i} tail (Verdes-Montenegro et al. 2006). We used SE's Kron elliptical apertures to measure magnitudes (mag$_{-}$auto) in FUV, NUV and R-bands. Although mag$_{-}$auto is often used to measure total magnitudes in galaxy surveys (e.g. Bell et al. 2004, de Mello et al. 2006, Zucca et al. 2006), high uncertainties might be expected at the faint magnitudes due to the assumption that the sky background has Gaussian random noise without source confusion (Brown et al. 2007). However, since we are comparing data taken with different resolutions and the fact that UV light does not necessarily peak at the same coordinate as the optical light, mag$_{-}$auto performs better than the other SExtractor choices, as long as a careful match between the different catalogs is performed. We matched the coordinates of the sources in the 3-bands within 1$''$ to 5$''$ and after visually checking each identification we chose 3$''$ as our best matched catalog with FUV, NUV and R-detections. The cross-match in the three bands resulted not only in the four most luminous members of HCG 100 (Table 1) but also 16 other objects, within the 10$'$$\\times$10$'$ field (Table 2). Nine of these sources are within the H{\\sc i} tail: \\# 3 is the most distant object from the group (381 arcsec or 137.2 kpc from HCG100a) and is located in the southern tip of the H{\\sc i} tail, \\#4 is far from all of the brightest members of the group and located in the densest H{\\sc i} region, \\#5 and \\#6 are close to HCG100c, \\#8 is isolated and 150 arcsec (53.9 kpc) from HCG100a, \\#9 is close to HCG100a, \\#13 is a small galaxy not originally classified as a member of the HCG100 and located in the optical tidal tail of HCG100b, \\#14 and \\#15 are also close to the tidal tail. The other six objects have no H{\\sc i} detection but are still within the chosen field. SE's magnitudes in FUV, NUV and R (mag$_{-}$auto) were corrected for foreground Galactic extinction using E(B-V)=0.081 and A$_{R}$=E(B-V)$\\times$2.634 (Schlegel et al. 1998), A$_{FUV}$=E(B-V)$\\times$8.29 and A$_{NUV}$=E(B-V)$\\times$8.18 (Seibert et al. 2005). In Table 3 we list the H{\\sc i} mass per beam in the vicinity of nine of the intragroup objects detected, the other 7 objects are below the detection limit ($<1.7 \\times 10^{8}$). However, it is important to note that, due to the low resolution of the data, the HI masses of the objects near large galaxies are contaminated by the HI masses of the latter. Only objects \\#3, \\#4 and \\#8 are far away from bright members of the group and this contamination could be avoided. Objects \\#13, \\#14, \\#15 besides being close to H100b are within the same beam, therefore the HI masses listed in Table 3 are not the individual masses of each object but the total mass within one beam centered in that region. As seen in Fig.\\ref{fuvhi_10arcmin} most of the intragroup objects are located in the outskirts of the H{\\sc i} contours where column densities range from 7.5 $\\times 10^{19}$ to 5$\\times 10^{20}$ cm$^{-2}$, except for object \\#4 which is located in a peak of H{\\sc i} where NH{\\sc i} is 1.2$\\times 10^{21}$ cm$^{-2}$. Therefore, H{\\sc i} clouds to which many of them are associated have column densities about one order of magnitude lower than the N(HI)$\\sim$10$^{21}$ cm$^{-2}$, value thought to be required for triggering star formation (e.g. Skillman et al. 1988). Due to the low resolution of the HI data, all values are lower limits to the true HI column density. \\begin{deluxetable}{cccccccc} \\tablecaption{HCG100 Properties} \\tablewidth{0pt} \\tablehead{ \\colhead{ID} & \\colhead{Morphology$^{\\rm a}$} & \\colhead{Velocity$^{\\rm b}$} & \\colhead{FUV$^{\\rm c}$} & \\colhead{NUV} & \\colhead{R$^{\\rm d}$} & \\colhead{FUV$_{\\rm corr}$$^{\\rm e}$} & \\colhead{NUV$_{\\rm corr}$} } \\startdata HCG100 a & Sb & 5323 &17.30 & 16.78 & 12.5 & 16.12 & 16.63\\\\ HCG100 b & Sm & 5163 &17.49 & 17.10 & 14.1 & 16.44 & 16.82\\\\ HCG100 c & SBc& 5418 &18.46 & 17.91 & 14.7 & 17.25 & 17.79\\\\ HCG100 d & Scd& &19.08 & 18.65 & 15.5 & 17.99 & 18.41\\\\ \\enddata \\tablenotetext{a}{Morphology from Plana et al. 2003} \\tablenotetext{b}{Systemic Velocity in kms$^{-1}$ from Plana et al. 2003} \\tablenotetext{c}{FUV and NUV magnitudes were obtained using IRAF task ellipse.} \\tablenotetext{d}{R-band magnitudes are from Rubin et al. 1991.} \\tablenotetext{e}{Extinction corrections using Seibert et al. (2005) for FUV and NUV.} \\end{deluxetable} \\begin{deluxetable}{ccccrrrr} \\tabletypesize{\\scriptsize} \\tablecaption{FUV sources within HCG100 $10' \\times 10'$ field} \\tablewidth{0pt} \\tablehead{ \\colhead{ID} & \\colhead{RA} & \\colhead{Dec} & \\colhead{R$^{\\rm a}$} & \\colhead{NUV} & \\colhead{FUV} & \\colhead{FUV-R} & \\colhead{FUV-NUV} } \\startdata 1 & 0.2410& 13.0647& 18.80 $\\pm$ 0.01& 20.56 $\\pm$ 0.04& 20.44 $\\pm$ 0.06& 1.64 $\\pm$ 0.06 & -0.12 $\\pm$ 0.07\\\\ 2 & 0.2526& 13.1161& 19.54 $\\pm$ 0.01& 21.61 $\\pm$ 0.08& 21.48 $\\pm$ 0.10& 1.94 $\\pm$ 0.10 & -0.13 $\\pm$ 0.13\\\\ 3 & 0.2722& 13.0236& 19.09 $\\pm$ 0.01& 21.61 $\\pm$ 0.08& 21.21 $\\pm$ 0.09& 2.12 $\\pm$ 0.09 & -0.40 $\\pm$ 0.12\\\\ 4 & 0.2929& 13.0859& 20.25 $\\pm$ 0.04& 21.26 $\\pm$ 0.07& 21.09 $\\pm$ 0.09& 0.84 $\\pm$ 0.10 & -0.17 $\\pm$ 0.12\\\\ 5 & 0.2933& 13.1377& 19.42 $\\pm$ 0.01& 22.43 $\\pm$ 0.15& 22.79 $\\pm$ 0.22& 3.37 $\\pm$ 0.22 & 0.36 $\\pm$ 0.27\\\\ 6 & 0.3076& 13.1630& 17.48 $\\pm$ 0.00& 20.19 $\\pm$ 0.03& 20.20 $\\pm$ 0.05& 2.72 $\\pm$ 0.05 & 0.01 $\\pm$ 0.06\\\\ 7 & 0.3176& 13.0290& 18.43 $\\pm$ 0.00& 22.85 $\\pm$ 0.11& 21.70 $\\pm$ 0.12& 3.27 $\\pm$ 0.12 & -1.15 $\\pm$ 0.16\\\\ 8 & 0.3188& 13.0724& 20.39 $\\pm$ 0.01& 22.49 $\\pm$ 0.13& 21.18 $\\pm$ 0.11& 0.79 $\\pm$ 0.11 & -1.31 $\\pm$ 0.17\\\\ 9 & 0.3429& 13.1208& 20.35 $\\pm$ 0.01& 22.24 $\\pm$ 0.11& 22.10 $\\pm$ 0.16& 1.75 $\\pm$ 0.16 & -0.15 $\\pm$ 0.19\\\\ 10& 0.3535& 13.0704& 19.75 $\\pm$ 0.01& 21.92 $\\pm$ 0.09& 21.66 $\\pm$ 0.11& 1.91 $\\pm$ 0.11 & -0.26 $\\pm$ 0.14\\\\ 11& 0.3627& 13.0532& 19.42 $\\pm$ 0.01& 21.63 $\\pm$ 0.07& 21.62 $\\pm$ 0.10& 2.20 $\\pm$ 0.10 & -0.01 $\\pm$ 0.12\\\\ 12& 0.3704& 13.1375& 18.88 $\\pm$ 0.00& 22.15 $\\pm$ 0.09& 21.98 $\\pm$ 0.14& 3.10 $\\pm$ 0.14 & -0.18 $\\pm$ 0.17\\\\ 13$^{\\rm b}$&\t 0.3735& 13.0986& 17.69 $\\pm$ 0.00&\t 19.63 $\\pm$ 0.02& 19.59 $\\pm$ 0.04& 1.90 $\\pm$ 0.04 & -0.04 $\\pm$ 0.04\\\\ 13A& 0.3738& 13.0987& 17.83 $\\pm$ 0.01& 19.78 $\\pm$ 0.05& 19.81 $\\pm$ 0.17& 1.98 $\\pm$ 0.17 & 0.03 $\\pm$ 0.18\\\\ 13B& 0.3719& 13.0984& 17.77 $\\pm$ 0.01& 19.62 $\\pm$ 0.04& 19.83 $\\pm$ 0.04& 2.06 $\\pm$ 0.04 & 0.22 $\\pm$ 0.06\\\\ 14& 0.3773& 13.1041& 20.42 $\\pm$ 0.01& 21.63 $\\pm$ 0.08& 21.10 $\\pm$ 0.09& 0.69 $\\pm$ 0.09 & -0.53 $\\pm$ 0.12\\\\ 15& 0.3796& 13.0950& 19.52 $\\pm$ 0.01& 22.04 $\\pm$ 0.12& 21.79 $\\pm$ 0.14& 2.27 $\\pm$ 0.14 & -0.25 $\\pm$ 0.19\\\\ 16& 0.3816& 13.0817& 18.89 $\\pm$ 0.00& 21.55 $\\pm$ 0.06& 21.30 $\\pm$ 0.11& 2.41 $\\pm$ 0.11 & -0.24 $\\pm$ 0.12\\\\ \\enddata \\tablenotetext{a}{Magnitudes (AB) in all bands were obtained with SExtractor Mag$_{-}$auto. Galactic extinction corrections were done using Schlegel et al. (1998) for R and Seibert et al. (2005) for FUV and NUV.} \\tablenotetext{b}{Object 13 was separated into two objects, A and B, using IRAF polyphot task.} \\end{deluxetable} \\begin{deluxetable}{cccccccccc} \\tabletypesize{\\scriptsize} \\tablecaption{Derived Properties} \\tablewidth{0pt} \\tablehead{ \\colhead{ID} & \\colhead{L$_{\\rm FUV}$ (erg/s)$^{\\rm a}$} & \\colhead{L$_{\\rm NUV}$ (erg/s)} & \\colhead{age$^{\\rm b}$} & \\colhead{SFR$_{\\rm FUV}$ $^{\\rm c}$} & \\colhead{SFR$_{\\rm NUV}$} & \\colhead{log(M$_{\\rm *}$)$^{\\rm d}$} & \\colhead{log(M$_{\\rm HI}$)$^{\\rm e}$} & \\colhead{Sep$^{\\rm f}$} & \\colhead{log I$_{1530}$$^{\\rm g}$} } \\startdata 1 & 5.90E+040 & 5.70E+040 & 3.9\t & 0.005& 0.007 & 5.0 & & 131.2\t & 6.02 \\\\ 2 & 2.27E+040 & 2.16E+040 & 3.8\t & 0.002& 0.003 & 4.6 & & 102.2\t & 5.91 \\\\ 3 & 2.91E+040 & 2.17E+040 & $<$1\t & 0.002& 0.003 & 4.7 & 9.6 & 137.2 & 5.90 \\\\ 4 & 3.25E+040 & 2.98E+040 & 3.3\t & 0.003& 0.004 & 4.7 & 10.4 & 60.7 & 5.35 \\\\ 5 & 6.79E+039 & 1.02E+040 & 194.1\t & 0.001& 0.001 & 6.5 & 9.9$\\dagger$ & 61.2 & 4.99 \\\\ 6 & 7.41E+040 & 8.01E+040 & 26.9\t & 0.006& 0.010 & 6.3 & 9.2$\\dagger$ & 74.7 & 6.31 \\\\ 7 & 1.84E+040 & 6.90E+039 & $<$1\t & 0.001& 0.001 & 4.5 & & 108.5\t & 5.51 \\\\ 8 & 2.98E+040 & 9.62E+039 & $<$1\t & 0.002& 0.001 & 4.7 & 10.0 & 53.9 & 5.31 \\\\ 9 & 1.28E+040 & 1.21E+040 & 3.5\t & 0.001& 0.001 & 4.6 & 9.8$\\dagger$ & 17.5 & 5.80 \\\\ 10 & 1.92E+040 & 1.63E+040 & 2.9\t & 0.002& 0.002 & 4.4 & & 58.8\t & 5.97 \\\\ 11 & 2.00E+040 & 2.13E+040 & 21.0\t & 0.002& 0.003 & 5.6 & & 83.7\t & 5.85 \\\\ 12 & 1.43E+040 & 1.31E+040 & 3.2\t & 0.001& 0.002 & 4.3 & & 58.0\t & 5.95 \\\\ 13 & 1.29E+041 & 1.35E+041 & 13.7\t & 0.010& 0.016 & 6.2 & 9.5$\\dagger$$\\dagger$ & 53.3 & 6.67 \\\\ 13 A& 1.03E+041 & 1.36E+041 & 118.7\t & 0.008& 0.017 & & & & 6.51 \\\\ 13 B& 1.06E+041 & 1.17E+041 & 33.8\t & 0.008& 0.014 & & & & 6.50 \\\\ 14 & 3.22E+040 & 2.13E+040 & $<$1\t & 0.003& 0.003 & 4.8 & 9.5$\\dagger$$\\dagger$ & 56.3\t& 6.30 \\\\ 15 & 1.70E+040 & 1.46E+040 & 2.9\t & 0.001& 0.002 & 4.3 & 9.5$\\dagger$$\\dagger$ & 62.1 & 5.07 \\\\ 16 & 2.67E+040 & 2.29E+040 & 2.9\t & 0.002& 0.003 & 4.5 & & 72.0\t & 5.26 \\\\ \\enddata \\tablenotetext{a}{FUV and NUV luminosities are in erg/s, divide by the FUV and NUV bandwidths (269\\AA\\ and 616\\AA, respectively) to obtain in L erg/s/\\AA.} \\tablenotetext{b}{Age (Myr) from FUV-NUV using Thilker et al. (2007) assuming a Milky Way internal extinction (E(B-V)=0.2).} \\tablenotetext{c}{SFR (M$_{\\odot}$/yr) from Iglesias-P\\'aramo et al. (2006) using FUV and NUV without correcting for internal extinction.} \\tablenotetext{d}{Stellar mass (M$_{\\odot}$) obtained from Starburst99 using ages (column 4) and L$_{\\rm 1530}$ (erg/s/\\AA).} \\tablenotetext{e}{MHI (M$_{\\odot}$) was calculated using 2.36 $\\times$ 10$^{5}$ F$_{\\rm HI}$ D$^{2}$, where D is in Mpc and F$_{\\rm HI}$ in Jy Km/s. F$_{\\rm HI}$ was measured within one beam size 61.0$''$ $\\times$ 55.23$''$ and reflects the HI mass in the vicinity of each object.} \\tablenotetext{f}{Distance (kpc) between each object and HCG100a.} \\tablenotetext{g}{FUV surface brightness (L$_{\\odot}$ kpc$^{-2}$) is defined following Hoopes at al. (2007).} \\tablenotetext{\\dagger}{HI masses of Objects \\#5, \\#6, and \\#9 should be taken with caution due to contamination by large galaxies in the vicinity.} \\tablenotetext{\\dagger\\dagger}{HI masses of Objects \\#13, \\#14, and \\#15 are not individual masses of each object but the total mass within one beam centered in that region. Contamination from H100b mass is also possible.} \\end{deluxetable} ", "conclusions": "We have analyzed the UV and optical light in combination with the H{\\sc i} gas within the compact group of galaxies HCG100 and identified 16 star-forming regions in the intragroup region. The young age ($<200$~Myr) of these objects and the proximity to the tidal tail connects the OB stars formation time scale ($\\sim$10$^{8}$ yr) with the dynamic time scale of the tidal features. Moreover, the H{\\sc i} clouds to which many of them are associated have column densities about one order of magnitude lower than the N(HI)$\\sim$10$^{21}$ cm$^{-2}$ thought to be required for triggering star formation. So, in these cases, we have a strong indication that the H{\\sc i} clouds must have suffered recent collisions which could have then triggered the star formation process. Based on their ages, stellar masses and H{\\sc i} masses in their vicinities, we conclude that some of these objects are tidal dwarf galaxies with ongoing star formation and some are intergalactic HII regions or conglomeration of stellar clusters. Although we have an estimate of the amount of neutral gas in these objects, the main ingredient in the star-formation process, the molecular gas, is unknown for these objects. From CO observations of tidal dwarf galaxies (TDGs), Braine et al. (2001) provided strong evidence that TDGs are self-gravitating entities with large amounts of atomic gas which will transform into molecular gas and subsequently form stars. CO observations together with spectroscopy of these objects will be of great value in understanding their nature. For instance, we cannot exclude the possibility that some of these objects are not associated with HCG100, but at least those falling within HI density peaks are likely to be associated with the group. Moreover, we will be able to estimate metallicity of these newly formed regions and evaluate the enrichment level of the processed gas in the intragroup medium. \\begin{deluxetable}{lcccc} \\tablecaption{Comparison of Galaxy Properties} \\tablewidth{0pt} \\tablehead{ \\colhead{Parameter$^{\\rm a}$} & \\colhead{Large UVLGs} & \\colhead{Compact UVLGs} & \\colhead{LBGs} & \\colhead{HCG100 IGs} } \\startdata log L$_{1530}$ (L$_{\\odot}$) & 10.3 to 11.2 & 10.3 to 11.0 & 10.3 to 10.9 & 8 to 9 \\\\ log I$_{1530}$ (L$_{\\odot}$ kpc$^{-2}$) $^{\\rm b}$& 6.0 to 8.0 & 8.0 to 10.3 & 9.0 to 10.3 & 5.6 to 8.3 \\\\ log SFR (M$_{\\odot}$ yr$^{-1}$) $^{\\rm c}$ & 0 to 1.5 & 0.2 to 2.0 & 0.5 to 2.0 & -2 to -0.4 \\\\ FUV-R $^{\\rm d}$ & 1.0 to 3.5 & 0.2 to 2.8 & 0.2 to 1.7 & 0.7 to 3.4 $^{\\rm f}$ \\\\ \\enddata \\tablenotetext{a}{Parameters for UVLGS and LBGs are from Hoopes et al. (2007). The lowest value for HCG100 intragroup objects was calculated using average values A$_{1500}$=1.6 and the highest value using A$_{1500}$=4.1 from Calzetti (2001).} \\tablenotetext{b}{I$_{1530}$ is the FUV surface brightness as described in Hoopes et al. (2007), except for HCG100 intragroup objects where we used R as the FUV petrosian radius.} \\tablenotetext{c}{SFR for the HCG intragroup objects using FUV equation from Iglesias-P\\'aramo et al. (2006).} \\tablenotetext{d}{r magnitudes are from SDSS and corrected for extinction, except for HCG100 which is the same as in Table 1 (Cousin R magnitude) and is not corrected for extinction.} \\tablenotetext{e}{FUV-R shown is not corrected for internal extinction. Using Calzetti (2001) A$_{1500}$=1.6 and A$_{r}$=0.48 FUV-R=-0.3 to 2.2.} \\end{deluxetable} \\clearpage \\begin{figure} \\plotone{f1.eps} \\caption{GALEX FUV image 10$'$$\\times$10$'$ with H{\\sc i} contours. Four HCG100 members are marked as `a, b, c and d'. Intragroup objects are circled (radius=8$''$) and numbered. VLA NH{\\sc i} contours are 0.6, 1.2, 2.1, 3.6, 4.4, 5.1, 5.9, 6.6, 7.4 $\\times$10$^{20}$ cm$^{-2}$, beam size (61.0$''$$\\times$55.23$''$) is shown on left corner. North is to the top and East to the left. \\label{fuvhi_10arcmin}} \\end{figure} \\begin{figure} \\plotone{f2.eps} \\caption{GALEX NUV image 10$'$$\\times$10$'$ with H{\\sc i} contours. Intragroup objects are circled (radius=8$''$) and numbered. VLA NH{\\sc i} contours are 0.6, 1.2, 2.1, 3.6, 4.4, 5.1, 5.9, 6.6, 7.4 $\\times$10$^{20}$ cm$^{-2}$. Beam size (61.0$''$$\\times$55.23$''$) is shown in Fig.\\ref{fuvhi_10arcmin}. North is to the top and East to the left. \\label{nuvhi_10arcmin}} \\end{figure} \\begin{figure} \\plotone{f3.eps} \\caption{GALEX FUV-NUV versus age from Thilker et al. models (2007) are shown as solid line (no extinction correction) and dotted line (internal extinction of the Milky Way E(B-V)=0.2). 4 objects are not included, they have ages $<$ 1~Myr old. \\label{colorsfuvnuvthilker}} \\end{figure} \\begin{figure} \\plotone{f4.eps} \\caption{GALEX FUV-NUV versus FUV-R of the intragroup objects. Squares and triangles are FUV--R colors assuming an error of 0.2 magnitudes in the R-magnitude. Models from Thilker et al. 2007 are shown as solid line (no extinction correction) and dotted line (internal extinction of the Milky Way E(B-V)=0.2). Crosses mark ages 10, 50, 100, 200, and 500 Myr, from the bottom left to the top right. \\label{colorsfuvnuvfuvr}} \\end{figure} \\begin{figure} \\plotone{f5.eps} \\caption{R images of all intragroup objects, circle has radius=8$''$, each cutout is 1$'$$\\times$1$'$. NH{\\sc i} contours (white) are the same as in Fig.\\ref{fuvhi_10arcmin}. \\label{all_obj}} \\end{figure} \\clearpage" }, "0710/0710.4397_arXiv.txt": { "abstract": "We describe a finite-volume method for solving the Poisson equation on oct-tree adaptive meshes using direct solvers for individual mesh blocks. The method is a modified version of the method presented by Huang and Greengard (2000), which works with finite-difference meshes and does not allow for shared boundaries between refined patches. Our algorithm is implemented within the FLASH code framework and makes use of the PARAMESH library, permitting efficient use of parallel computers. We describe the algorithm and present test results that demonstrate its accuracy. ", "introduction": "\\label{Sec:intro} Astrophysical simulations commonly need to solve the Poisson equation, \\begin{equation} \\label{Eqn:Poisson} \\nabla^2 \\phi({\\bf x}) = 4 \\pi G \\rho({\\bf x})\\ , \\end{equation} for the gravitational potential $\\phi({\\bf x})$ given a density distribution $\\rho({\\bf x})$. Similar equations also arise in other contexts, such as incompressible flow problems and divergence-cleaning methods for magnetohydrodynamics. Self-gravitating problems offer special challenges because they frequently develop structure spanning large spatial dynamic ranges. The problem of spatial dynamic range is particularly acute for grid-based schemes for solving the Euler equations of hydrodynamics. Within the context of grid-based methods for solving the Poisson equation, several approaches to the problem of spatial dynamic range have arisen. The simplest approach is to use Fourier transforms, multigrid methods, or sparse iterative solvers on uniform Eulerian grids. The maximum dynamic range is then limited by the available memory. Recently Trac and Pen (2006) have demonstrated an out-of-core uniform-grid Poisson solver that exceeds this limit by making use of disk space; the largest published calculations with this solver have used $4000^3$ zones. However, storage resource consumption still increases with the third power of the resolution, putting grids with $10^4$ zones on a side or larger out of reach for now. If high resolution is not needed everywhere in the domain, as is frequently the case in cosmological structure formation simulations, it is also possible to employ nonuniform Eulerian or Lagrangian grids. Examples include COSMOS (Ricker, Dodelson, \\& Lamb 2000), which uses a nonuniform multigrid solver, and MMH (Pen 1998), which uses a deformable mesh. These methods work best when the region to be resolved is known beforehand, although fully Lagrangian codes like Pen's can follow the development of structures and adjust zone spacing appropriately. Nonuniform grids, however, introduce complicated position-dependent stencils and generally cannot be used with fast transform-based solvers. In addition, coupled numerical hydrodynamics methods generally place constraints on the allowed mesh anisotropy and nonuniformity, since numerical dissipation increases with zone spacing. The greatest spatial dynamic ranges in grid-based astrophysical simulations have been achieved using adaptive mesh refinement (AMR) techniques. Modern AMR techniques for solving hyperbolic systems of equations were first developed by Berger and Oliger (1984) and Berger and Colella (1989). In the Berger and Colella formulation, AMR involves the construction of a hierarchical set of mesh ``patches'' with decreasing zone spacing. The coarsest mesh covers the entire computational domain, while more highly refined meshes cover only a portion. Generally refined meshes are taken to be nested; that is, each refined mesh lies completely within its coarser parent mesh. Examples of astrophysical codes employing patch-based AMR meshes include the code of Truelove et al.\\ (1998), AMRA (Plewa \\& M\\\"uller 2001), RIEMANN (Balsara 2001), Enzo (O'Shea et al.\\ 2004), and CHARM (Miniati \\& Colella 2007). To date self-gravitating AMR calculations have achieved effective spatial resolutions greater than $10^{15}$. A considerable simplification of the Berger and Colella method was introduced by Quirk (1991) and de~Zeeuw and Powell (1993). Known as ``oct-tree'' AMR, this method requires that each refined patch contain the same number of zones, that each refinement level have zones a factor of two smaller in each dimension than the next coarser level, and that each refined patch be no more than one level removed from its immediate neighbor. Mesh data can then be stored in an oct-tree data structure, allowing for extremely efficient parallel implementations, even on high-latency systems (Warren \\& Salmon 1993). Also, because each mesh patch (often called a ``block'' in this context) contains the same number of zones, it is possible to achieve high levels of cache re-use when iterating over zones. Unless each block contains a very small number of zones, this efficiency comes with the price that refined blocks often must cover more of the volume than they would in a patch-based method. The primary astrophysical simulation code employing oct-tree AMR is FLASH (Fryxell et al.\\ 2000), which uses the PARAMESH library (MacNeice et al.\\ 2000) to handle its AMR mesh. (The ART (Kravtsov, Klypin, \\& Khoklov 1997), MLAPM (Knebe, Green, \\& Binney 2001), and RAMSES (Teyssier 2002) codes also use tree structures to manage AMR meshes, but in these codes the refined blocks consist of a single zone each, and the base mesh generally contains a large number of zones, so these codes do not employ oct-trees. A block-based AMR approach that allows for ``incomplete families'' has also recently been implemented within the VAC code (van der Holst \\& Keppens 2007); the oct-trees discussed in the current paper require complete ``families'' of child blocks.) By default PARAMESH uses blocks containing $8^3$ interior zones as a compromise between adaptive flexibility and memory efficiency, but any size larger than the differencing stencil and small enough to fit in the memory attached to a single processor can be employed. Poisson solvers on oct-tree meshes generally employ some type of multigrid or sparse linear solver iteration scheme. Transform methods cannot be employed directly because of the varying mesh resolution and non-tensor-product character of the composite mesh. For example, Matsumoto and Hanawa (2003) describe a relaxation-based method for solving the Poisson equation on nested grids. Within the FLASH framework, we have employed the Martin and Cartwright (1996) multigrid algorithm for several years. However, the speed and scalability of this algorithm have been limited because of the need to apply multiple relaxation iterations on each level, together with the communication of block boundary data that such iterations require. Because the cost of this algorithm dominates the cost of most self-gravitating simulations with FLASH, we are motivated to develop more efficient methods that require less communication. In this paper we describe one such method, based on the direct multigrid algorithm of Huang and Greengard (2000, hereafter HG). This algorithm improves considerably upon relaxation-based multigrid solvers by allowing refined patches to be solved directly using ``black-box'' uniform-grid solvers. Unlike the direct method described by Couchman (1991), the HG algorithm properly minimizes the global residual by allowing information to flow back from fine meshes to coarse meshes. However, it is formulated on a finite-difference mesh in which refined patches are not permitted to touch. Here we describe a modified version of the HG algorithm suitable for finite-volume oct-tree AMR meshes. We have implemented this algorithm within the FLASH framework, and we present test results that demonstrate the solver's accuracy. The present paper should be regarded as a companion to Fryxell et al.\\ (2000) and a methodological description for future FLASH-based simulation papers in the areas of cosmic structure formation, galaxy cluster physics, star cluster formation, and binary star evolution, among others. The paper is organized as follows. In \\S~\\ref{Sec:algorithm} we give a precise description of the algorithm. In \\S~\\ref{Sec:tests} we present the results of test problems run with the new solver. We conclude in \\S~\\ref{Sec:conclusions} with some remarks on performance. All calculations described in this paper were performed using version 2.4 of FLASH. ", "conclusions": "\\label{Sec:conclusions} We have detailed the modifications needed in order to use the Huang \\& Greengard (2000) algorithm for Poisson's equation on oct-tree AMR meshes. This method allows us to use a local direct Poisson solver with Dirichlet boundary conditions on each block, yet it correctly minimizes the residual across the composite AMR mesh. The HG algorithm must be modified to take into account the fact that mesh quantities represent zone-averaged values and the fact that block boundaries can coincide on an oct-tree mesh. Because adjacent coarse and fine blocks do not share points in common for oct-tree meshes, additional interpolation (as compared to HG) is needed to set boundary values. The resulting errors at block corners reduce convergence to first order, but a fixed small number of boundary-zone relaxation steps restores the desired second-order convergence rate for block corners in uniformly refined regions. A higher-order scheme may be able to eliminate the need for such relaxation. Our test results show that, while jumps in refinement degrade convergence somewhat in comparison with solving on a uniform mesh, the effect is manageable because oct-tree meshes are usually fully refined up to some level. Thus first-order convergence takes over only once the error has been significantly reduced at second order on fully refined levels. The parallel scaling of this solver, as implemented using the Message Passing Interface (MPI) in the FLASH code, is comparable to or slightly better than that of the relaxation multigrid solver distributed with FLASH 2.$x$. With constant total work and increasing processor count, parallel efficiency is close to 100\\% up to 8 -- 16 times the smallest number of processors on which a run can fit. When gasdynamics is included, the Poisson solver requires $\\sim$50\\% of the execution time; for particle-only simulations the amount is $\\sim $70 -- 80\\%. In comparison with the older solver, a factor of 2 -- 5 improvement in performance is often seen. The solver described here will be made available to the public as part of FLASH~3.0.\\footnote{FLASH is freely available at http://flash.uchicago.edu/.} The primary scaling bottleneck for this and other multigrid algorithms is the fact that on the coarsest level there are too few zones to distribute among all of the processors. Since each V-cycle descends to the coarsest level, processors controlling blocks on finer levels must wait until the coarsest level is finished before proceeding. To counteract this work starvation, we are investigating the use of a uniform-grid parallel FFT solver (T.~Theuns, private communication) to handle the coarsest level in a distributed fashion. Additional strategies may be necessary to make optimal use of the petaflop computing resources now beginning to become available." }, "0710/0710.4674_arXiv.txt": { "abstract": "We present a study of large-scale bars in the local Universe, based on a large sample of $\\sim3692$ galaxies, with $-18.5\\leq M_g < -22.0$ mag and redshift $0.01\\leq z<0.03$, drawn from the Sloan Digitized Sky Survey. While most studies of bars in the local Universe have been based on relatively small samples that are dominated by bright galaxies of early to intermediate Hubble types with prominent bulges, the present sample is $\\sim$~10 times larger, covers a larger volume, and includes many galaxies that are disk-dominated and of late Hubble types. Both color cuts and S\\'ersic cuts yield a similar sample of $\\sim2000$ disk galaxies. We characterize bars and disks by ellipse-fitting $r$-band images and applying quantitative criteria. After excluding highly inclined ($>60^{\\circ}$) systems, we find the following results. (1)~The optical $r$-band fraction ($f_{\\rm opt-r}$) of barred galaxies, when averaged over the whole sample, is $\\sim48\\%-52\\%$. The bars have diameters $d$ of 2 to 24 kpc, with most ($\\sim72\\%$) having $d\\sim$ 2 to 6 kpc. (2)~When galaxies are separated according to half light radius ($r_{\\rm e}$), or normalized $r_{\\rm e}$/$R_{\\rm 24}$, which is a measure of the bulge-to-disk ($B/D$) ratio, a remarkable result is seen: $f_{\\rm opt-r}$ rises sharply, from $\\sim$~40\\% in galaxies that have small $r_{\\rm e}$/$R_{\\rm 24}$ and visually appear to host prominent bulges, to $\\sim$~70\\% for galaxies that have large $r_{\\rm e}$/$R_{\\rm 24}$ and appear disk-dominated. Visual classification, performed for $\\sim900$ galaxies, confirms our result that disk-dominated galaxies with no bulge or a very low $B/D$ display a significantly higher optical bar fraction ($>70$\\% vs 40\\%) than galaxies with prominent bulges. It also shows that barred galaxies host a larger fraction (31\\% vs 5\\%) of quasi-bulgeless disk-dominated galaxies than do unbarred galaxies. (3)~$f_{\\rm opt-r}$ rises for galaxies with bluer colors (by $\\sim30\\%$) and lower masses (by $\\sim15\\%-20\\%$). (4) The significant rise in the optical bar fraction toward late-type galaxies is discussed in terms of their higher gas mass fraction, higher dark matter fraction, and lower bulge-to-disk ratio. (5) While hierarchical $\\Lambda$CDM models of galaxy evolution models fail to produce galaxies without classical bulges, our study finds that $\\sim20\\%$ of disk galaxies appear to be ``quasi-bulgeless''. (6)~Our study of bars at $z\\sim0$ in the optical $r$-band provides the $z\\sim0$ comparison point for $HST$ ACS surveys (e.g., GEMS, GOODS, COSMOS) that measure the rest-frame optical bar fraction in bright galaxies out to $z\\sim1$. After applying the same cutoffs in magnitude, bar ellipticity ($e_{\\rm bar}\\ge0.4$), and bar size ($a_{\\rm bar}\\ge1.5$ kpc), which are applied in $z\\sim0.2-1.0$ studies in order to trace strong bars with adequate spatial resolution in bright disks, we obtain an optical $r$-band bar fraction of $34\\%$. This is comparable to the value reported at $z\\sim0.2-1.0$, implying that the optical fraction of strong bars does not suffer a dramatic order of magnitude decline in bright galaxies out to $z\\sim1$. ", "introduction": "The majority ($\\sim$~60\\%) of bright disk galaxies are barred, when observed in the near-infrared \\citep[][hereafter MJ07]{kna99,esk00,lau04,men07,mar07} and a significant fraction of these ($\\sim$~45\\%) also appear barred in the optical \\citep[][MJ07]{esk00}. Earlier studies suggested a striking or order of magnitude decline in the optical fraction of bars out to $z\\sim1$ \\citep{abr99,vdb00}, but subsequent studies have ruled out an order of magnitude decline and find that the optical fraction of strong bars remains fairly constant or show a moderate decline of a factor of $\\sim2$ \\citep[][Sheth et al. 2003,2004,2007; see $\\S$ \\ref{discu}]{jog04,elm04a,zhe05}. Bars are believed to be very important with regard to the dynamical and secular evolution of disk galaxies, particularly in redistributing the angular momentum of the baryonic and dark matter components of disk galaxies \\citep{com81,wei85,com90,deb00}. The interaction between the bar and the disk material can lead to the inflow of gas from the outer disk to the central parts, which can trigger starbursts \\citep{elm94,kna95,reg04,jog05,sht05} and might contribute to the formation of disky bulges \\citep[or 'pseudobulges',][]{kor93,kor04,ath05a,jog05,deb06}. Additional evidence for secular evolution is provided by box- or peanut-shaped bulges in inclined galaxies. These features are commonly attributed to the orbital structure, resonances, and vertical instabilities in a barred potential \\citep{com90,kui95,bur99,bur05}. From a theoretical perspective, it is possible to model some aspects of the evolution of disks and bars, and their interactions (e.g., the corresponding simulations are able to reproduce certain broad features of barred disks). However, it remains unclear why a specific galaxy has a bar, but a seemingly similar galaxy is unbarred; or why some barred galaxies have a classical bulge, whereas others harbor a disky bulge, etc. This might indicate that specific properties of the disks or the particular processes involved in their formation have a strong impact on their ability to form a bar. In order to investigate, how disk and bar formation are related, it is not only important to determine the fraction of disk galaxies that are barred, but also to relate bar and disk properties. There are different methods to find and characterize bars. The Third Reference Catalog of Bright Galaxies \\citep[][hereafter RC3]{dev91} uses three bar strength families (SA, SAB, and SB) to characterize bars based on a visual inspection of blue light images. Using this classification \\cite{ode96} showed that the optical fraction of strong bars in disk galaxies rises from Sc galaxies towards later-types. More quantitative measures, such as the gravitational torque method \\citep{blo02,lau02,but05}, or Fourier dissection \\citep{but06,lau06}, were also used, not only to find bars, but to quantitatively determine bar strengths and bar lengths. Similarly, the method of fitting ellipses to galaxy isophotes provides a tool to characterize the length and shape of bars \\citep{fri96,jog99,kna00,sht00,lai02,why02,jog02a,jog02b,sht03,elm04a,ree07,mar07,men07,sht07}. These efforts were able to shed light on the fraction, shapes, and structures of bars in local disk galaxies of early to intermediate Hubble types. First attempts were made to relate the presence of a bar or its structural properties to other galaxy characteristics. However, there were three important limitations. Firstly, samples used in earlier studies were small ($\\sim100$ to 200 objects) and mostly composed of bright galaxies of early to intermediate Hubble types (Sa to Sc), with fairly prominent bulges. One could barely get decent number statistics for bars in early-type disk galaxies, while the bins of disk-dominated late Hubble types were dominated by Poisson noise (e.g., see Figure 16 in MJ07). Secondly, with such small samples, it was difficult to bin galaxies in terms of the galaxy host properties. Thirdly, earlier samples were drawn from a very small volume, and could be highly impacted by cosmic variance. In the present study, we use a sample of $\\sim2000$ galaxies, at $z=0.01-0.03$ with $M_r \\sim -18.5$ to $-22.0$ mag. The first advantage of this study is that it provides a factor of 10 improvement in number statistics and reduces the effect of cosmic variance by selecting galaxies drawn from a larger volume. Secondly, with $\\sim2000$ galaxies, we can for the first time have $100-200$ galaxies per bin, while binning galaxies in terms of host galaxy parameters, such as luminosity, measures of bulge-to-disk ($B/D$) ratios, size, colors, surface brightness, etc. This allows us to conduct a comprehensive study of barred and unbarred galaxies as a function of host galaxy properties. Thirdly, our sample has a large number of galaxies, which are relatively faint ($M_g>-19.5$ mag) or/and appear disk-dominated, characteristic of late Hubble types. This allows us to shed light on what happens to bars at the fainter end of the luminosity function and in the regime of disk-dominated galaxies. A fourth goal of our study is to provide a reference baseline for bars at $z\\sim0$ in the {\\it rest-frame optical} for intermediate redshift $HST$ surveys using the Advanced Camera for Surveys (ACS), such as the Tadpole field \\citep{tra03}, the Galaxy Evolution from Morphologies and SEDs \\citep[GEMS,][]{rix04}, the Great Observatories Origins Deep Survey \\citep[GOODS,][]{gia04}, and COSMOS \\citep{sco06}, which trace bars in the rest-frame optical band at $z\\sim0.2-1.0$ (look-back times of 3--8 Gyr). We use SDSS to provide the reference point at $z=0$ in the $r$-band, complementing the one in the $B$-band of MJ07. Our $B$- and $r$-band results can be directly compared to $HST$ ACS optical studies of bars in bright disks at $z\\sim0.2-1.0$ \\citep{elm04a,jog04}. The validity of this comparison is reinforced by the fact that we use the same procedure of ellipse fits ($\\S$ \\ref{meth}) that were used by these studies. We also note that the reference $z=0$ point for bars in the near-infrared band \\citep{men07} is not appropriate for comparison with the above $HST$ ACS surveys, which trace the rest-frame optical rather than the rest-frame near-infrared. The outline of the paper is as follows: In $\\S$ \\ref{samsel} we present our sample selection. The method used to find and characterize bars is explained in $\\S$ \\ref{meth}. In $\\S$ \\ref{resol} we discuss the detection limits. Our results and more detailed assessments of specific findings are presented in $\\S$ \\ref{resul}. We discuss our results in $\\S$ \\ref{discu} and summarize our conclusions in $\\S$ \\ref{sum}. Throughout the paper we assume a flat cosmology with $\\Omega_M=1-\\Omega_{\\Lambda}=0.3$ and $H_0=70$ km~s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "\\label{sum} We have used the $r$-band images from the NYU-VAGC of a sample of 3692 galaxies with $-18.5\\leq M_g < -22.0$ mag and redshift $0.01\\leq z<0.03$ to find and characterize bars. While most studies of bars in the local Universe have been based on relatively small samples that are dominated by bright early type (Sa to Sc) galaxies with bulges, the present sample also includes many galaxies that are disk-dominated and of late Hubble types. Furthermore, the sample is $\\sim$~10 times larger and samples a larger volume than earlier local samples We used a color cut in the color-magnitude diagram to select $\\sim2000$ disk galaxies. We cross-check that S\\'ersic cuts would yield a similar sample. We identify and characterize bars and disks using $r$-band images and a method based on ellipse fits and quantitative criteria. The typical seeing ($1\\farcs4$ or 290 to 840 pc over $0.01\\leq z<0.03$) is adequate for resolving large-scale bars, whose typical diameters are $\\ge$ 2 kpc. Smaller nuclear bars are not the focus of this study. After the standard procedure of excluding highly inclined ($>60^{\\circ}$) systems, we find the following results. \\begin{enumerate} \\item The average optical $r$-band bar fraction ($f_{\\rm opt-r}$) in our sample, which primarily consists of late-type disk-dominated galaxies, is $\\sim48\\%-52\\%$. The bars have diameters $d$ of 2 to 24 kpc, with most ($\\sim72\\%$) having $d\\sim$ 2 to 6 kpc (Figure \\ref{barp}a). The bar length is typically much smaller than $R_{24}$ (Figure \\ref{r24}a) and most galaxies have a $a_{bar}/R_{24}$ in the range 0.2 to 0.4 (Figure \\ref{r24}b). \\item When galaxies are separated according to normalized $r_{\\rm e}$/$R_{\\rm 24}$, which is a measure of the bulge-to-disk ($B/D$) ratio, a remarkable result is seen: the optical $r$-band fraction rises sharply, from $\\sim$~40\\% in galaxies that have small $r_{\\rm e}$/$R_{\\rm 24}$ and visually appear bulge-dominated, to $\\sim$~70\\% for galaxies that have large $r_{\\rm e}$/$R_{\\rm 24}$. Visual classification of $\\sim80\\%$ of our sample (with $i<60^{\\circ}$) confirms our result that {\\it late-type disk-dominated galaxies with no bulge or a very low $B/D$ display a significantly higher optical bar fraction ($>70$\\% vs 40\\%) than galaxies with prominent bulges.} It also shows that barred galaxies host a larger fraction (31\\% vs 5\\%) of quasi-bulgeless disk-dominated galaxies than do unbarred galaxies. The bar ellipticities or strengths are on average higher in faint disk-dominated galaxies than in bulge-dominated galaxies (Figure \\ref{visp2}d). \\item Similar trends in the optical bar fraction are found using the central surface brightness and color. Bluer galaxies have higher bar fractions ($\\sim58\\%$ at $g-r=0.3$) than the redder objects ($\\sim32\\%$ at $g-r=0.65$) (Figure \\ref{barf}d). The optical $r$-band fraction also shows a slight rise for galaxies with fainter luminosities (Figure \\ref{barf}c) and lower masses (Figure \\ref{massp}). This is expected from (2), given that late-type galaxies are fainter, bluer, and less massive. \\item The significant rise in the optical bar fraction toward disk-dominated galaxies is discussed in terms of their higher gas mass fraction, higher dark matter fraction, and lower bulge-to-disk ratio. \\item While many hierarchical $\\Lambda$CDM models of galaxy evolution models fail to produce galaxies without classical bulges, our study finds that in the range $-18.5\\leq M_g < -22.0$ mag and redshift $0.01\\leq z<0.03$, $\\sim20\\%$ of the 1144 moderately inclined disk galaxies appear to be ``quasi-bulgeless'', without a classical bulge. \\item Our study of bars at $z\\sim0$ in the optical $r$ band provides a reference $z\\sim0$ baseline for intermediate redshift $HST$ ACS surveys that trace bars in {\\it bright} disks in the rest-frame optical bands ($BVRI$) out to $z\\sim1$. By applying the same cutoffs in magnitude, bar ellipticity ($e_{\\rm bar} \\geq0.4$), and bar size ($a_{\\rm bar} \\geq1.5$ kpc), which are applied in $z\\sim0.2-1.0$ studies in order to trace strong bars with adequate spatial resolution in bright disks, we obtain an optical $r$-band fraction for strong bars of $34\\%$. This is comparable to the values of $\\sim30\\%$ at $z\\sim0.2-1.0$, $\\sim$~36\\%~$\\pm$~6\\% at $z\\sim0.2-0.7$, and $\\sim$~24\\%~$\\pm$~4\\% at $z\\sim0.7-1.0$. Our result implies that the optical fraction of strong bars in bright galaxies does not suffer any dramatic order of magnitude decline out to $z\\sim$~1. \\end{enumerate}" }, "0710/0710.4442_arXiv.txt": { "abstract": "We construct for the first time, the sequences of stable neutron star (NS) models capable of explaining simultaneously, the glitch healing parameters, $Q$, of both the pulsars, the Crab ($Q \\geq 0.7$) and the Vela ($Q \\leq 0.2$), on the basis of starquake mechanism of glitch generation, whereas the conventional NS models cannot give such consistent explanation. Furthermore, our models also yield an upper bound on NS masses similar to those obtained in the literature for a variety of modern equations of state (EOSs) compatible with causality and dynamical stability. If the lower limit of the observational constraint of (i) $Q \\geq 0.7$ for the Crab pulsar and (ii) the recent value of the moment of inertia for the Crab pulsar (evaluated on the basis of time-dependent acceleration model of the Crab Nebula) , $I_{\\rm Crab,45} \\geq 1.93$ (where $I_{45}=I/10^{45}\\,{\\rm g.cm}^2$), both are imposed together on our models, the models yield the value of matching density, $E_b = 9.584 \\times 10^{14}{\\rm\\,g\\,cm}^{-3}$ at the core-envelope boundary. This value of matching density yields a model-independent upper bound on neutron star masses, $M_{\\rm max} \\leq 2.22 M_\\odot$, and the strong lower bounds on surface redshift $z_R \\simeq 0.6232$ and mass $M \\simeq 2.11 M_\\odot$ for the Crab ($Q \\simeq 0.7$) and the strong upper bound on surface redshift $z_R \\simeq 0.2016 $, mass $M \\simeq 0.982 M_\\odot$ and the moment of inertia $I_{\\rm Vela,45} \\simeq 0.587$ for the Vela ($Q \\simeq 0.2$) pulsar. However, for the observational constraint of the `central' weighted mean value $Q \\approx 0.72$, and $I_{\\rm Crab,45} > 1.93$, for the Crab pulsar, the minimum surface redshift and mass of the Crab pulsar are slightly increased to the values $z_R \\simeq 0.655$ and $M \\simeq 2.149 M_\\odot$ respectively, whereas corresponding to the `central' weighted mean value $Q \\approx 0.12$ for the Vela pulsar, the maximum surface redshift, mass and the moment of inertia for the Vela pulsar are slightly decreased to the values $z_R \\simeq 0.1645,\\, M \\simeq 0.828 M_\\odot$ and $I_{\\rm Vela,45} \\simeq 0.459$ respectively. These results set an upper and lower bound on the energy of a gravitationally redshifted electron-positron annihilation line in the range of about 0.309 - 0.315 MeV from the Crab and in the range of about 0.425 - 0.439 MeV from the Vela pulsar. ", "introduction": "The data on the glitch healing parameter, $Q$, provide the best tool for testing the starquake (Ruderman 1972; Alpar et al 1996) and Vortex unpinning (Alpar et al 1993) models of glitch generation in pulsars. Both of these mechanisms of glitch generation, in fact, consider NSs, in general, a two component structure: a superfluid interior core surrounded by a rigid crust (in the present study we shall use the term `envelope' which includes the solid crust and other interior portion of the star right up to the superfluid core). In the starquake model, $Q$ is defined as the fractional moment of inertia, i.e. the ratio of the moment of inertia of the superfluid core, $I_{\\rm core}$, to the moment of inertia of the entire configuration, $I_{\\rm total}$, as (Pines et al 1974) \\begin{equation} Q = \\frac{I_{\\rm core}}{I_{\\rm total}}. \\end{equation} Recently, Crawford \\& Demia\\'{n}ski (2003) have collected the all measured values of the glitch healing parameter $Q$ for Crab and Vela pulsars available in the literature and found that for 21 measured values of $Q$ for Crab glitches, a weighted mean of the values yields $Q = 0.72 \\pm 0.05$, and the range of $Q \\geq 0.7$ encompasses the observed distribution for the Crab pulsar. In order to test the starquake model for the Crab pulsar, they have computed $Q$ (as given by Eq.(1)) values for seven representative EOSs of dense nuclear matter, covering a range of neutron star masses. Their study shows that the much larger values of $Q(\\geq 0.7)$ for the Crab pulsar is fulfilled by all the six EOSs (out of seven considered in the study) corresponding to a `realistic' neutron star mass range $1.4\\pm 0.2M_\\odot$. By contrast, a weighted mean value of the 11 measurements for Vela yields a much smaller value of $Q(= 0.12 \\pm 0.07)$ and the all estimates for Vela agree with the likely range of $Q \\leq 0.2$. Thus, their results are found to be consistent with the starquake model predictions for the Crab pulsar. They have also concluded that the much smaller values of $Q \\leq 0.2$ for the Vela pulsar are inconsistent with the starquake model predictions, since the implied Vela mass based upon their models corresponds to a value $ \\leq 0.5M_\\odot$ for $Q \\leq 0.2$, which is too low as compared to the `realistic' NS mass range. Thus, in the literature, the starquake is considered as a viable mechanism for glitch generation in the Crab and the vortex unpinning, the another mechanism, is considered suitable for the Vela pulsar, since it can avoid some other problems associated with the starquake explanation of the Vela glitches (see, e.g. Crawford \\& Demia\\'{n}ski (2003); and references therein). However, it seems surprising that if the internal structure of NSs are described by the same two component conventional models (as mentioned above), different kinds of glitch mechanisms are required for the explanation of a glitch! Furthermore, it also follows from the above discussion that the main reason for not considering the starquake, the feasible mechanism for glitch generation in the Vela, lies in the fact that there exists none of the sequence of NS models in the literature which could explain simultaneously, on the basis of starquake model, both the extreme limiting cases of glitch healing parameter, $Q$, corresponding to the Vela ($Q \\leq 0.2$) and the Crab ($Q \\geq 0.7$) pulsars in the range $0 \\leq Q \\leq 1.0$ for the `realistic' NS mass values for both the pulsars. The present study, therefore, deals with the construction of such models \\footnote{however, the other problems associated with the starquake explanation of the Vela glitches (see, e.g. Crawford \\& Demia\\'{n}ski (2003); and references therein) are not considered in the present paper. The future study in this regard may provide some explanation, provided the correlation between various parameters of the Crab and the Vela pulsar, obtained in the present study, can be utilized.} We assume that {\\em all} the NSs belong to the same family of NS sequence which terminates at the {\\em maximum} value of mass. Certainly, this {\\em maxima} should correspond to an {\\em upper bound} on NS masses. In order to construct such a sequence, we have to set the extreme causal EOS (in geometrized units), $dP/dE = 1$ (where $P$ is the pressure and $E$ the energy-density) to describe the core. Firstly, because various observational studies like - the gamma-ray burst data, X-ray burst data and the glitch data etc., and their explanation (see, e.g. Lindblom 1984; Cottam et al 2002; Datta \\& Alpar 1993) favour the stiffest EOSs. The latest estimate of the moment of inertia for the Crab pulsar (based upon the `newest' observational data on the Crab nebula mass) rules out most of the existing EOSs of the dense nuclear matter, leaving only the stiffest ones (Bejger \\& Haensel 2002; Haensel et al 2006). Secondly, because of the fact that the `real' EOS of the dense nuclear matter beyond the density range $\\sim 10^{14}\\,{\\rm g\\, cm}^{-3}$ are largely unknown due to the lack of knowledge of nuclear interactions (see, e.g. Dolan 1992; and references therein; Haensel et al 2006), and the various EOSs available in the literature (see, e.g. Arnett \\& Bowers 1977) for NS matter represent only an extrapolation of the results far beyond this density range. Though, the status of the `real' EOS for NS matter is not certain, one could impose some well-known physical principle, independent of the EOS, such as the `causality condition' ($dP/dE = 1$) throughout the core of the star beyond a fiduciary density, $E_b$, at the core-envelope boundary to ascertain a definite upper bound on NS masses (see, e.g., Rhoades \\& Ruffini 1974; Hartle 1978; Lindblom 1984; Friedman \\& Ipser 1987; Kalogera \\& Baym 1996). In this connection this is also to be pointed out here that the maximum mass for {\\em any} EOS describing the core, beyond the density $E_b$, with a subluminal sound velocity turns out to be less than that of the upper bound obtained by using the extreme causal EOS (see, e.g., Haensel et al 2006). The envelope of our models (below the density $E_b$ at the core-envelope boundary) may be characterized by the well-known EOS of classical polytrope ${\\rm d}$ln$P/{\\rm d}$ln$\\rho = \\Gamma_1$ (where $\\rho$ denotes the density of the rest-mass and $\\Gamma_1$ is a constant known as the adiabatic index) for different values of the constant $\\Gamma_1 = (4/3), (5/3)$ and 2 respectively. The reason for considering the polytropic EOS for the entire envelope lies in the fact that with this EOS, our models yield an upper bound on NS masses {\\em independent} of the value of $\\Gamma_1$, and this upper bound (for a fiduciary choice of $E_b$) is found fully consistent with those of the values cited in the literature (Kalogera \\& Baym 1996; Friedman \\& Ipser 1987). Thus, the choice of the said polytropic EOS for the entire envelope may be regarded entirely equivalent to the choice of the various EOSs like WFF (Wiringa, Fiks \\& Fabrocini 1988), FPS (Lorenz, Ravenhall \\& Pethick, 1993), NV (Negele \\& Vautherin 1973), or BPS (Baym, Pethick \\& Sutherland 1971) in an appropriate sequence below the density range $E_b$, adopted by various authors in the conventional models of NSs (see, e.g., Kalogera \\& Baym 1996; Friedman \\& Ipser 1987), so that the constant $\\Gamma_1$ appearing in the polytropic EOS may be looked upon as an `average' $\\Gamma_1$ for the density range below $E_b$, specified by the sequence of various EOSs in the conventional models of NSs. The choice of the constant $\\Gamma_1 = 4/3, 5/3$ and 2 thus become obvious, since this choice can cover almost the entire range of density discussed in the literature for NS matter which is also applicable for the envelope region - the polytropic EOS with $\\Gamma_1 = 4/3$ represents the EOS of extreme relativistic degenerate electrons and non-relativistic nuclei (Chandrasekhar 1935), $\\Gamma_1 = (5/3)$ represents the well-known EOS of non-relativistic degenerate `neutron gas' (Oppenheimer \\& Volkoff 1939), and $\\Gamma_1 = 2$ represents the case of extreme relativistic baryons interacting through a vector meson field (Zeldovich 1962) (The value of $\\Gamma_1 > 2$ is also possible for some EOS describing the NS matter, e.g., Malone, Johnson \\& Bethe 1975; Clark, Heintzmann \\& Grewing 1971, however, the results obtained in this paper remain unaffected for the choice of $\\Gamma_1 > 2$), and the outcome of this study (in terms of explaining the glitch healing parameter for various pulsars and predicting the upper bound on the compactness of NSs (since the upper bound on mass is independent of the value of $\\Gamma_1$)) would finally decide, among the chosen values, the `appropriate' value of $\\Gamma_1$ for the NS envelope. The validity of assuming the extreme causal EOS in the core and a polytropic EOS in the envelope of the present models, in view of the various modern EOS of dense nuclear matter, is also discussed in the last section of the present paper. We have noted that in all conservative models of NSs, the choice of the core-envelope boundary, $r_b$ (corresponding to a density denoted by $E_b$), is somewhat {\\em arbitrary} in the sense that there are no criteria available for the choice of a particular matching density, $E_b$, below which the EOS of the NS matter is assumed to be known and unique. One can freely choose somewhat lower values of $E_b$ (which will increase the core size) to obtain higher values of $Q$ (see, e.g. Shapiro \\& Teukolsky 1983; Datta \\& Alpar 1993). To avoid such a procedure, we choose the core-envelope boundary of our models on the basis of the `compatibility criterion' which asserts that for an assigned value of the ratio ($\\sigma$) of central pressure, $P_0$, to central energy-density, $E_0$, the compactness parameter $u(\\equiv M/R$; total mass to radius ratio in geometrized units) of any {\\em regular} configuration should not exceed the compactness parameter $u_h$ of the homogeneous density sphere, in order to assure the compatibility with the hydrostatic equilibrium (Negi \\& Durgapal 2001; Negi 2004a). This criterion is capable of constraining the core-envelope boundary of any physically realistic NS model. A combination of this criterion with those of the observational data on the glitch healing parameter and the recently estimated minimum value of the moment of inertia for the Crab pulsar (based on the newly estimated `central value' of the Crab nebula mass $M {\\rm (nebula)} \\simeq 4.6 M_\\odot$ in the time-dependent acceleration model), $I_{\\rm Crab,45} = 1.93$; where $I_{\\rm Crab,45} = I_{\\rm Crab}/10^{45}$ g\\,cm$^2$ (Bejger \\& Haensel 2003) can provide the desired NS models discussed above, since both the theory (criterion) and the observations (stated above) are being used to construct the NS models. ", "conclusions": "This study constructs the stable sequences of NS models terminate at the value of maximum mass, $M_{\\rm max} \\simeq 2.22 M_\\odot$, independent of the EOSs of the envelope, for the matching density, $E_b = 9.584 \\times 10^{14}$ g\\, cm$^{-3}$, at the core-envelope boundary. This value of `matching density' is a consequence of the observational constraints $Q \\simeq 0.7$ and $I_{\\rm Crab,45} \\simeq 1.93$ (associated with the Crab pulsar) imposed together on the $\\Gamma_1 = 2$ envelope model and in this sense does not represent a fiduciary quantity. The upper bound of the surface redshift, $z_R \\simeq 0.77$ (corresponding to a $u$ value $\\simeq 0.34$), however, belongs to the model with a $\\Gamma_1 = 2$ envelope which is consistent with the absolute upper bound on the surface redshift of NS models compatible with causality and pulsational stability (Negi 2004b). This special feature, together with some other remarkable ones, discussed in the present study underline the appropriateness of the $\\Gamma_1 = 2$ envelope model. Since among the variety of modern EOSs discussed in the literature, the upper bound on NS mass compatible with causality and dynamical stability can reach a value up to $2.2M_\\odot$ (in this category, the SLy (Douchin \\& Haensel 2001) EOS yields a maximum mass of $2.05M_\\odot$, whereas the BGN1 (Balberg \\& Gal 1997) and the APR (Akmal et. al. 1998) EOSs yield the maximum masses of $2.18M_\\odot$ and $2.21M_\\odot$ respectively (see, e.g. Haensel et al 2006)). In view of this result, the model-independent maximum mass, $M_{\\rm max} \\simeq 2.22M_\\odot$, obtained in this study may be regarded as good as those obtained on the basis of modern nuclear theory. In addition to this result, the appropriate sequences of stable NS models obtained in this study can explain the glitch healing parameter, $Q$, of any glitching pulsar, provided the weighted mean values of $Q$ lie in the range $0 < Q \\leq 0.78$. This finding also reveals that if the starquake is considered to be a viable mechanism for glitch generation in all pulsars, then the envelope of `real' NSs may be well approximated by a polytropic EOS corresponding to a polytropic index, $n$, closer to 1. For the value of matching density, $E_b = 9.584 \\times 10^{14}$ g\\, cm$^{-3}$, the $\\Gamma_1 = 2$ envelope model yields the minimum values of mass $M \\simeq 2.11 M_\\odot$ and surface redshift $z_R \\simeq 0.6232$ for the Crab ($Q \\simeq 0.7$) and the maximum values of mass $M \\simeq 0.982 M_\\odot$ and surface redshift $z_R \\simeq 0.2016$ for the Vela pulsar ($Q \\simeq 0.2$). The minimum mass and surface redshift for the Crab pulsar are slightly increased up to the values $M \\simeq 2.149 M_\\odot$ and $z_R \\simeq 0.655$ respectively, if the `central' weighted mean value of $Q \\approx 0.72$ and the moment of inertia $I_{\\rm Crab,45} > 1.93 $ are also imposed on these models. However, for the `central' weighted mean value of $Q \\simeq 0.12$ corresponding to the Vela pulsar, the maximum mass and surface redshift are somewhat decreased to the values $M \\simeq 0.828 M_\\odot$ and $z_R \\simeq 0.1645$ respectively. This value of mass and surface redshift for the Vela pulsar can further decrease up to the values $M \\simeq 0.685 M_\\odot$ and $z_R \\simeq 0.1312$ respectively, if the lower weighted mean value of $Q \\simeq 0.05$ for the Vela pulsar is imposed. These results predict the upper and lower bounds on the energy of a gravitationally redshifted electron-positron annihilation line in the range of about 0.309 - 0.315 MeV from the Crab and in the range of about 0.425 - 0.439 MeV from the Vela pulsar respectively. For a comparison, if the observational constraint of the minimum value of $I_{\\rm Crab,45} \\simeq 3.04$ (the value of moment of inertia for the Crab pulsar obtained earlier by Bejger \\& Haensel (2002), on the basis of the constant-acceleration model for the Crab nebula) together with $Q \\simeq 0.7$ is imposed on the models studied in the present paper, the $\\Gamma_1 = 2$ envelope model yields the value of matching density, $E_b = 7.0794 \\times 10^{14}$ g\\, cm$^{-3}$. This value of $E_b$ yields a model-independent upper bound on NS mass $M_{\\rm max} \\simeq 2.59 M_\\odot$. This value of maximum mass, however, represents an `average' of the maximum NS masses in the range $2.2M_\\odot \\leq M_{\\rm max} \\leq 2.9M_\\odot$ obtained by Kalogera \\& Baym 1996 (and references therein) on the basis of other EOSs for NS matter, fitted to experimental nucleon-nucleon scattering data and the properties of light nuclei. For this lower value of matching density, the $\\Gamma_1 = 2$ envelope models yield the minimum value of mass $M \\simeq 2.455 M_\\odot$ for the Crab ($Q \\simeq 0.7$) and the maximum value of mass $M \\simeq 1.142 M_\\odot$ for the Vela pulsar ($Q \\simeq 0.2$). The minimum mass for the Crab pulsar is slightly increased up to the value $M \\simeq 2.5 M_\\odot$, if the `central' weighted mean value of $Q \\approx 0.72$ and the moment of inertia $I_{\\rm Crab,45} > 3.04 $ are also imposed on these models. However, corresponding to the `central' weighted mean value of $Q \\simeq 0.12$, the maximum mass of the Vela pulsar is somewhat decreased to the value $M \\simeq 0.964 M_\\odot$. This value of mass for the Vela pulsar can further decrease up to the value $M \\simeq 0.796 M_\\odot$, if the lower weighted mean value of $Q \\simeq 0.05$ for the Vela pulsar is imposed. Furthermore, the study can also explain some special features associated with the extraordinary gamma-ray burst of 5 March 1979." }, "0710/0710.4168_arXiv.txt": { "abstract": "We have applied the torus fitting procedure described in Ng \\& Romani (2004) to PWNe observations in the \\emph{Chandra} data archive. This study provides quantitative measurement of the PWN geometry and we characterize the uncertainties in the fits, with statistical errors coming from the fit uncertainties and systematic errors estimated by varying the assumed fitting model. The symmetry axis $\\Psi$ of the PWN are generally well determined, and highly model-independent. We often derive a robust value for the spin inclination $\\zeta$. We briefly discuss the utility of these results in comparison with new radio and high energy pulse measurements. ", "introduction": "One of the greatest success of the Chandra X-ray Observatory (CXO) is the discovery of equatorial tori and polar jet structures in many pulsar wind nebula (PWN) systems. It is now believe that these features are common among young neutron stars. In the \\citet{ree74} picture, when the highly relativistic pulsar wind decelerates in the external medium, a termination shock is formed at a characteristic scale \\[r_t =\\left(\\frac{\\dot E}{4\\pi c \\eta P_{\\rm ext}} \\right )^{1/2} \\ , \\] where $\\dot E$ is the pulsar spin-down power and $\\eta$ is the filling factor. In general, if the pulsar is subsonic in the ambient medium, i.e. $ P_{\\rm ext} \\geq P_{\\rm ram}= 6\\times 10^{-10}nv^2_7\\;\\mathrm {g\\;cm^{-1}s^{-2}}$ for a pulsar speed $10^7v_7\\;\\mathrm{cm\\;s^{-1}}$ through a density of $n\\,m_p\\;\\mathrm{cm^{-3}}$, a toroidal shock structure is expected; faster objects produce bow shock nebulae. Indeed, many young pulsars still inside their high pressure supernova remnant birth sites do show such toroidal symmetry. The best-known example is the PWN around the \\object{Crab pulsar} as observed by the \\emph{CXO} \\citep{wei01}. Recently, several relativistic MHD models, e.g.\\ \\citet{kom03} and \\citet{del06}, have shown how such toroidal structure can form if the pulsar wind has a latitudinal variation. \\citet{ng04} (hereafter \\citetalias{ng04}) developed a fitting procedure to measure the 3D orientation of the pulsar wind torus and applied to a few X-ray observations. While this simple geometrical model does not capture the fine details of the MHD simulations, is does allow one to extract the torus (and hence pulsar spin) orientation from relatively low signal-to-noise ratio (S/N) data. \\citetalias{ng04} also gave quantitative estimates for the statistical errors arising from Poisson statistics. However the systematic errors due to unmodeled components such as jets or background were neglected. For bright objects e.g.\\ the Crab and Vela pulsars, the S/N is high and such systematic errors dominate. In this study, we attempt to quantify these systematic errors and apply the fitting to more \\emph{CXO} PWN observations, thus providing a more comprehensive study. ", "conclusions": "In conclusion, we have applied the torus fitting technique to more PWN observations in the \\emph{Chandra} data archive and characterized the uncertainties in the fits. This study provides a better understanding of the systematic errors, giving quantitative estimates of the measurement uncertainties. We argue that these robust position angle $\\Psi$ and inclination $\\zeta$ values are particularly useful for comparison with the radio and high energy pulse data. If new observations can fill in more measurements from these energy bands in Table 3, we should be able to make substantial progress in understanding the emission zones and viewing geometries of young pulsars." }, "0710/0710.1860_arXiv.txt": { "abstract": "We present a comprehensive analysis of structure in the young, embedded cluster, NGC~1333 using members identified with {\\it Spitzer} and 2MASS photometry based on their IR-excess emission. In total, 137 members are identified in this way, composed of 39 protostars and 98 more evolved pre-main sequence stars with disks. Of the latter class, four are transition/debris disk candidates. The fraction of exposed pre-main sequence stars with disks is $83\\% \\pm 11\\%$, showing that there is a measurable diskless pre-main sequence population. The sources in each of the Class~I and Class~II evolutionary states are shown to have very different spatial distributions relative to the distribution of the dense gas in their natal cloud. However, the distribution of nearest neighbor spacings among these two groups of sources are found to be quite similar, with a strong peak at spacings of 0.045~pc. Radial and azimuthal density profiles and surface density maps computed from the identified YSOs show that NGC~1333 is elongated and not strongly centrally concentrated, confirming previous claims in the literature. We interpret these new results as signs of a low velocity dispersion, extremely young cluster that is not in virial equilibrium. ", "introduction": "Observations of embedded, star-forming clusters and groups show that the stellar distributions are often elongated, clumpy, or both \\citep[cf.][]{ll03,gute05,teix06,alle07}, and the structure seems tied to the distribution of dense gas in the clusters' natal molecular clouds. Most clouds have some non-spherical structure, yet current results suggest that the associated clusters have varying degrees of agreement with the cloud's structure depending on how deeply the members are embedded \\citep{gute05,teix06}. Given the high frequency of asymmetric structure in clouds, it seems reasonable to assume that the exposed, relatively structureless clusters we observe may have been asymmetric in a previous epoch. With the ejection of the majority of their natal gas and adequate time to migrate from their birth sites, the imprint of the underlying cloud structure could be lost rather quickly in a recently exposed young cluster. Thus the structure we measure in a distribution of young stellar objects (YSO) relative to the dense gas in the associated cloud may be a reasonable proxy for the current dynamical state of an embedded cluster. The launch of {\\it Spitzer} has provided a potent new facility for studies of the structure of young and embedded clusters. While some members of these young clusters are diskless pre-main sequence stars (Class~III), the majority of the membership are sources with excess emission at mid-IR wavelengths \\citep{hll01}, made up of a mixture of protostars (Class~I), still embedded and accreting from dense spherical envelopes, and the slightly more evolved pre-main sequence stars with circumstellar disks (Class~II). The mid-IR spectral energy distributions (SED) of these objects are dominated by the emission from their dusty circumstellar material, making them easily distinguishable from pure photospheric sources such as unrelated field stars and indistinguishable diskless cluster members. {\\it Spitzer}'s sensitivity to mid-IR emission makes it the best tool currently available for identifying {\\it and characterizing} YSOs with IR-excess, and that sensitivity is sufficient to detect these sources down to the Hydrogen--burning mass limit for regions within the nearest kiloparsec \\citep{gute04}. The complete census of YSOs with disks in a young cluster represents a high-confidence sample of bona fide cluster members, and for the youngest regions, such a sample includes a high fraction of the total number of members \\citep{hll01}. Furthermore, by separating the two canonical evolutionary classes of YSOs that {\\it Spitzer} so effectively detects, we are able to probe both the recent overall star-forming activity in the region as traced by the stars with disks, and the immediate star formation as represented by the protostars, since this phase is expected to be short-lived relative to the former. Many young clusters are dominated by heavy and spatially variable extinction from dust within their natal molecular cloud environment. Another advantage {\\it Spitzer} brings to studies of these regions is that the mid-IR wavelengths targeted by {\\it Spitzer}'s imaging instruments (Infrared Array Camera, or IRAC, at 3.6-8.0~$\\mu$m and Multiband Imaging Photometer for {\\it Spitzer}, or MIPS, at 24-160~$\\mu$m) are less affected by extinction from dust in comparison to near-IR or visible wavelengths \\citep[e.g.][]{ccm89}. Recent papers have done an excellent job of characterizing the reddening law in the IRAC bandpasses \\citep{inde05,flah07,huar07} and have made a first attempt at doing the same for the 24~$\\mu$m channel of MIPS \\citep{flah07,huar07}. Given the wealth of information {\\it Spitzer} can bring to the study of embedded clusters, we have surveyed over thirty clustered star-forming regions within the nearest kiloparsec through the Guaranteed Time Observations (GTO) program for the IRAC and MIPS instrument teams. With these data, we can not only provide a full census of the YSOs with IR-excess in each region, but also perform a detailed examination and comparative analysis of structure in young, embedded clusters. NGC~1333 has been a popular target for observations of deeply embedded protostars via radio wavelengths due to the presence of an unprecendented number of molecular outflows \\citep[e.g.][]{ks00} associated with several bright IRAS sources considered Class~0 protostars \\citep{jenn87}, all in relatively close proximity to the Sun \\citep[250~pc,][]{enoc06}. Many of the outflows are traced by shock-induced emission, a clear sign that they are indeed affecting the local, quiescent cloud material \\citep[e.g.][]{wala05}. Because of this, some studies have claimed that the NGC~1333 dense molecular cloud core is in the process of being destroyed by influence from outflows \\citep[e.g.][]{wari96,sk01,quil05}. In addition to the numerous protostars, there is a cluster of pre-main sequence stars identified by number counts analysis in the near-IR \\citep{svs76,asr94,lal96,wilk04}. \\citet{lal96} suggested that the distribution was well-described as a ``double cluster'', having two distinct surface density maxima. Here we present a {\\it Spitzer} IRAC and MIPS imaging and photometric analysis of the NGC~1333 young cluster, one of the most nearby large membership ($N>100$) clusters in the {\\it Spitzer} Young Cluster Survey. We achieve a near-complete census of the cluster membership that possesses circumstellar material, significantly surpassing the sensitivity of ground-based mid-IR surveys \\citep[cf.][and references therein.]{rebu03}. In addition, we statistically infer the population of pre-main sequence stars that lack disks, enabling an estimate of the fraction of members with disks. Finally, using the identified YSOs, we apply both established and recently developed methods for characterizing the structure of this cluster, in specific reference to previous claims of structure found in the literature. ", "conclusions": "We have employed the mid-IR sensitivity of {\\it Spitzer} to achieve a census of the YSO members of the NGC~1333 embedded cluster that was previously unattainable. Furthermore, we have shown that the sources identified in our {\\it Spitzer} census represent a large fraction of the total cluster membership (83\\%). With this penetrating view of the cluster, we have performed several measurements of different aspects of the structure of the cluster that are not confused by field star contamination and are less biased by extinction effects than previous studies. We confirm the double-peaked surface density morphology of the cluster reported in previous work \\citep{lal96}, with the caveat that this morphology is traced only by the more evolved Class~II population. The two main density peaks are located in local minima of the gas density distribution. In contrast, the protostars trace the dense gas in this region closely, as do a fraction of the Class~II. That gas distribution appears to be a network of filamentary structures, connecting the two density peaks into a single, albeit complex, structure. Despite the difference in spatial distributions, however, the Class~II and protostars have similarities in their nearest neighbor distance distributions, particularly in the location of the peak at spacings of 0.045~pc. These two evolutionary states are expected to differ in duration by an order of magnitude, thus a low overall velocity dispersion in the cluster and a very young Class~II population are needed to account for the similarity remaining in the spacings among the two populations. A further implication of the lack of evidence for dynamical evolution of the cluster is the need for the dispersal of the molecular gas by the Class~II sources on very short timescales. Given that the more evolved YSOs have not moved significantly from their birthsites, regions dominated by more evolved sources where gas densities are preferentially low, such as the double-peaks in the stellar distribution here, were once locations of high gas density. Thus it seems plausible that groups of low-mass stars are able to disrupt their dense natal gas {\\it locally}. In this sense, we concur with previous work that has suggested that the active outflows in NGC~1333 are destroying the cloud \\citep{sk01,quil05}. With a virtual lack of low or high mass stars and dense gas in the center of the ring-like structure of the cloud and cluster core, we argue that this is in fact a primordial structure and not outflow-evacuated. Considering all sources regardless of evolutionary class, the cluster is clearly elongated, an expected result given the double-peaked nature of the stellar surface density distribution and embedded nature of the cluster as a whole \\citep{gute05}. We have presented radial density profiles measured via several different methods, and all suggest a roughly uniform density distribution within a 0.3~pc radius, with a steep decline ($\\alpha=-3.3$) beyond. The flat central distribution and sharp radial decline in the surface density profile at radii larger than 0.3~pc also suggests a rather limited amount of dynamical interaction in this cluster. If we analyze this within the framework of dynamics-generated structures like those of King or EFF models, we have to choose extreme fitting parameters just to get marginal agreement with the measured profiles. The radial profile we have measured here has two clear regimes were a simple power law matchs the density profile well, and the sharp knee transition between them is poorly matched by either the King or EFF profiles. Furthermore, given the asymmetry of the NGC~1333 cluster and the presense of a considerable amount of structured, dense gas, the dynamical information inferred from either of these models is unlikely to be accurate. Cloud geometry is the most tenable cause for the structure we have observed. As such, the relative motions of the stars must be fairly slow and the Class~II population must be quite young in order to have preserved that structure into the Class~II evolutionary phase." }, "0710/0710.3054_arXiv.txt": { "abstract": "We report on the spatial relationship between solar flares and coronal mass ejections (CMEs) observed during 1996-2005 inclusive. We identified 496 flare-CME pairs considering limb flares (distance from central meridian $\\ge 45^\\circ$) with soft X-ray flare size $\\ge$ C3 level. The CMEs were detected by the Large Angle and Spectrometric Coronagraph (LASCO) on board the {\\it Solar and Heliospheric Observatory} ({\\it SOHO}). We investigated the flare positions with respect to the CME span for the events with X-class, M-class, and C-class flares separately. It is found that the most frequent flare site is at the center of the CME span for all the three classes, but that frequency is different for the different classes. Many X-class flares often lie at the center of the associated CME, while C-class flares widely spread to the outside of the CME span. The former is different from previous studies, which concluded that no preferred flare site exists. We compared our result with the previous studies and conclude that the long-term LASCO observation enabled us to obtain the detailed spatial relation between flares and CMEs. Our finding calls for a closer flare-CME relationship and supports eruption models typified by the CSHKP magnetic reconnection model. ", "introduction": "A solar flare is sudden flash of electromagnetic radiation (suggesting plasma heating) in the solar atmosphere, and a coronal mass ejection (CME) is an eruption of the atmospheric plasma into interplanetary space. Both phenomena are thought to be different manifestations of the same process which releases magnetic free energy stored in the solar atmosphere. The spatial relation between flares and CMEs contains information on the magnetic field configurations involved in the eruptive process and hence is important for modeling them. Many flare-CME models are based on the CSHKP (Carmichael, Sturrock, Hirayama, Kopp \\& Pneuman) magnetic reconnection model. The model requires that a flare occurs just underneath of an erupting filament which eventually becomes the core of the CME associated with the flare. Normally the core corresponds to the center of the CME, thus the CSHKP model requires that the flare occurs near the center of the CME span. Full-scale studies on the flare-CME relationship started in the 70s and 80s with the CME observations obtained by the {\\it Solwind} coronagraph on board {\\it P78-1} and the Coronagraph/Polarimeter telescope on board the {\\it Solar Maximum Mission} ({\\it SMM}). \\citet{harri86} carried out a detailed analysis of three flare-CME events observed by {\\it SMM} and reported that flares occurred near one foot of an X-ray arch, which is supposed to become a CME. He also analyzed 48 flare-CME events observed by {\\it SMM} and {\\it Solwind} and reported that many flares occurred near one leg of the associated CMEs. This result, called the flare-ejection asymmetry, is inconsistent with the CSHKP flare-CME model. \\citet{kahle89} examined 35 events observed by the {\\it Solwind} and reported that flare positions did not peak neither at the center nor at one leg of the CMEs. They concurred with Harrison at the point that the observations do not match with the CSHKP model, while disagreeing with the result that flares are likely to occur at one leg of CMEs. They pointed out that the parameter employed by Harrison was biased, and concluded that both observations are compatible with the fact that there is no preferred flare site with respect to the CME span. It should be noted that the two studies applied different criteria for the event selection. Harrison did not apply any criteria on flare X-ray intensity, flare location, and CME span, while Kahler et al. used only strong limb flares ($\\ge$ M1 level; central meridian distance (CMD) $\\ge 40^\\circ$) and wide CMEs (angular span $\\ge 40^\\circ$). Different criteria might produce different spatial distributions, but the results in both the studies were inconsistent with the schematic view of the CSHKP type flare-CME model. The Large Angle and Spectrometric Coronagraph \\citep[LASCO;][]{bruec95} on board the {\\it Solar and Heliospheric Observatory} ({\\it SOHO}) mission has observed more than 11,000 CMEs from 1996, which provides a great opportunity to investigate the flare-CME relationship. \\citet{harri06} reviewed several flare-CME studies and stated that \"the pre-SOHO conclusions about relative flare-CME locations and asymmetry are consistent with many recent studies.\" However, systematic statistical study is needed before reaching a firm conclusion. In this paper we revisit this issue using the large CME data obtained by {\\it SOHO} LASCO. \\begin{figure*} \\epsscale{0.90} \\plotone{f1.eps} \\caption{Three CMEs observed by {\\it SOHO} LASCO to illustrate the measurement of CME span. The top row shows direct images used to measure the main CME body, and the bottom row shows corresponding running difference images used to measure the whole CMEs. $\\phi_1$ and $\\phi_2$ indicate the PAs of side edges of the main CME body, and $\\phi_A$ and $\\phi_B$ indicate those of the whole CME. Arrows point to the position of the flares associated with the CMEs.} \\end{figure*} ", "conclusions": "" }, "0710/0710.1098_arXiv.txt": { "abstract": "% A planetary transit produces both a photometric signal and a spectroscopic signal. Precise observations of the transit light curve reveal the planetary radius and allow a search for timing anomalies caused by satellites or additional planets. Precise measurements of the stellar Doppler shift throughout a transit (the Rossiter-McLaughlin effect) place a lower bound on the stellar obliquity, which may be indicative of the planet's migration history. I review recent results of the Transit Light Curve project, and of a parallel effort to measure the Rossiter effect for many of the known transiting planets. ", "introduction": "% I have great admiration for the people who discover transiting planets. Identifying the candidate transit signals from among a hundred thousand light curves, and flushing out the numerous astrophysical false positives, are impressive feats. This article, however, is not about transit discovery, but rather about the next step: performing high-precision photometry and spectroscopy of exoplanetary transits. The goal of this step is to determine the planetary and stellar properties well enough to allow for meaningful comparisons with the familiar properties of the Solar system, and to inform our theories of planet formation. The most immediate result of transit photometry is a measurement of the planetary radius. In combination with the planetary mass, which can be inferred from the Doppler orbit of the star, these data give the first clues about the composition, interior structure, and atmospheric energy balance of the planet. An accurate radius is also needed to interpret the results of other observations, such as the detection of thermal emission or reflected light based on secondary-transit photometry. The timings of the transits can be used to refine the measurement of the orbital period and search for additional bodies in the system. In \\S~2, I describe the Transit Light Curve (TLC) project, an effort to gather high-precision photometry during exoplanetary transits. The most prominent spectroscopic signal during a transit is the Rossiter-McLaughlin (RM) effect. This effect is an anomalous Doppler shift that arises from stellar rotation. Measuring this effect allows one to assess the alignment between the planetary orbital axis and the stellar spin axis, a fundamental system property that provides clues about the process of planet migration. I describe some recent measurements of the RM effect in \\S~3. \\begin{figure}[!h] \\plotone{Lightcurves.eps} \\caption{{\\bf The Transit Light Curve project.} Each panel shows time-binned photometry from a campaign (2-5 separate transits) for a particular planet. Most of the data are $z$ band observations with the FLWO 1.2m telescope and Keplercam detector.} \\end{figure} ", "conclusions": "" }, "0710/0710.0866_arXiv.txt": { "abstract": "We discuss oxygen and iron abundance patterns in K and M red-giant members of the Galactic bulge and in the young and massive M-type stars inhabiting the very center of the Milky Way. The abundance results from the different bulge studies in the literature, both in the optical and the infrared, indicate that the [O/Fe]-[Fe/H] relation in the bulge does not follow the disk relation, with [O/Fe] values falling above those of the disk. Based on these elevated values of [O/Fe] extending to large Fe abundances, it is suggested that the bulge underwent a rapid chemical enrichment with perhaps a top-heavy initial mass function. The Galactic Center stars reveal a nearly uniform and slightly elevated (relative to solar) iron abundance for a studied sample which is composed of 10 red giants and supergiants. Perhaps of more significance is the fact that the young Galactic Center M-type stars show abundance patterns that are reminiscent of those observed for the bulge population and contain enhanced abundance ratios of $\\alpha$-elements relative to either the Sun or Milky Way disk at near-solar metallicities. ", "introduction": "Abundance patterns in different populations in the Milky Way can shed light on Galaxy formation and its chemical evolution. Studies carried out over the last decade have provided accurate abundance patterns for stars in the Milky Way disk and halo so that these populations are now fairly-well mapped. There is significantly less information, however, on the Galactic bulge population and, in particular the Galactic Center, due to difficulties associated with heavy extinction. The stellar content belonging to the bulge population is old, with an age of 10--12 Gyr (e.g. Zoccali et al. 2003) and chemical abundance studies in both low and high-resolution find that the bulge is overall metal-rich with a large abundance spread that spans $\\sim$1.5 dex (e.g. Fulbright et al. 2006). The Galactic Center, on the other hand, contains a significant young population, including many luminous and massive stars. Those known to date are concentrated in three clusters within 60 pc of the Galactic Center: the Central Cluster, the Arches Cluster and the Quintuplet Cluster. The youngest stars in these clusters were formed recently with ages ranging roughly between 1 -- 9 Myr. In this contribution we summarize the abundance results obtained recently for the bulge and Galactic center and briefly discuss their implications in light of Galactic chemical evolution. ", "conclusions": "Abundance patterns are discussed in two distinct populations: $\\sim$ 12 Gyr old red-giants from the bulge and the young and massive M-type giants and supergiants residing mostly within 2 pcs of the central Black Hole. The abundance results indicate that the iron abundance distribution for the admittedly small sample of Galactic Center targets studied so far shows very little abundance spread in contrast to the large metallicity spread found for the old bulge population. Most importantly, both populations show enhancements in the $\\alpha$-element abundances relative to the Galactic disk trend that might be explained invoking a top-heavy IMF." }, "0710/0710.2114_arXiv.txt": { "abstract": "{ We present new determination of the birth rate of AXPs and SGRS and their associated SNRs. We find a high birth rate of $1/(500\\ {\\rm yr})$ to $1/(1000\\ {\\rm yr})$ for AXPs/SGRs and their associated SNRs. These high rates suggest that all massive stars (greater than $\\sim (23$-$32) M_{\\odot}$) give rise to remnants with magnetar-like fields. Observations indicate a limited fraction of high magnetic fields in these progenitors thus our study necessarily implies magnetic field amplification. Dynamo mechanisms during the birth of the neutron stars require spin rates much faster than either observations or theory indicate. Here, we propose that neutron stars form with normal ($\\sim 10^{12}$ G) magnetic fields, which are then amplified to $10^{14}$-$10^{15}$ G after a delay of hundreds of years. The amplification is speculated to be a consequence of color ferromagnetism and to occur after the neutron star core reaches quark-deconfinement density. This delayed amplification alleviates many difficulties in interpreting simultaneously the high birth rate, high magnetic fields, and state of isolation of AXPs/SGRs and their link to massive stars. ", "introduction": "Early studies of association of Anomalous X-ray Pulsars (AXPs) with supernova remnants (SNRs) suggested that 5\\% of core-collapse SN results in AXPs (Gaensler et al. 1999). This was based on 3 SNR associations out of a total of 6 AXPs. Since then evidence has mounted that AXPs and soft gamma-ray repeaters (SGRs) are the same type of objects (Gavriil et al. 2002) and more AXPs, SGRs and associated SNRs have been identified. Thus it it timely to revisit the issue of AXPs/SGRs birthrates. In this study we present an updated investigation of the birth rate of AXPs/SGRs and in addition, for the first time, the birth rate of associated SNRs is given. Since AXPs/SGRs ages rely on spin-down age estimates whereas SNRs ages are based on shock expansion models, this constitutes two independent estimates for birth rates. Both samples yield a high birth rate for AXPs/SGRs\\footnote{An independent study by Gill\\&Heyl (2007), based on a population synthesis of AXPs detected in the ROSAT All-Sky Survey, yields a birth rate of $\\sim 0.22$ per century.} of $(1/5)$-$(1/10)$ of all core-collapse SNe, higher than previously appreciated. This high frequency of occurrence of AXPs/SGRs brings into focus issues related to the origin of the strong magnetic fields which we address here. This paper is presented as follows: \\S 2 describes the methods and presents the birth rate results, and \\S 3 discusses the implications. Our model, based on a delayed amplification of magnetic field, is presented in \\S 4 before concluding in \\S 5. ", "conclusions": "Our study of the birth rate of AXPs and SGRS and their associated SNRs suggest that about $1/5$ to $1/10$ of all core-collapse SN lead to AXPs/SGRs. These high rates suggest that all massive stars (greater than $ M_{\\rm low}$) give rise to remnants with magnetar-like fields. This raises these issues: (i) how do all progenitors with $M\\ge M_{\\rm low}$ generate $>10^{14}$ G fields in their compact remnants?; (ii) why is there a dichotomy in magnetic field strength between compact remnants from progenitors with mass greater than $M_{\\rm low}$ (i.e. $B\\sim 10^{14}$ G) and those with mass less than $\\sim M_{\\rm low}$ ($B\\sim 10^{12}$ G); (iii) why are all AXPs/SGRs isolated while many progenitors with $M>M_{\\rm low}$ are in binaries? In this study, we introduce the notion of delayed magnetic field amplification to resolve these issues. We propose that neutron stars from progenitor masses $M> 9M_{\\odot}$ are born with normal ($\\sim 10^{12}$ G) magnetic fields. A neutron star from a progenitor with an approximate mass range $M_{\\rm low}< M <60 M_{\\odot}$ will experience an explosive transition to a quark star (the QN) in which its magnetic field is amplified to $10^{14}$-$10^{15}$ G by color ferromagnetism (Iwazaki 2005). The second explosion (QN) and related mass loss helps to reduce the surviving compact binary fraction thus explaining the state of isolation of AXPs/SGRs. The transition occurs with a delay of several hundred years (Staff et al. 2006). This delayed amplification alleviates many difficulties in interpreting simultaneously the high birth rate and high magnetic fields of AXPs/SGRs and their link to massive stars." }, "0710/0710.0920_arXiv.txt": { "abstract": "Electric currents $j$ flow along the open magnetic field lines from the polar caps of neutron stars. Activity of a polar cap depends on the ratio $\\alph=j/c\\rhoGJ$, where $\\rhoGJ$ is the corotation charge density. The customary assumption $\\alph\\approx 1$ is not supported by recent simulations of pulsar magnetospheres and we study polar caps with arbitrary $\\alph$. We argue that no significant activity is generated on field lines with $0<\\alph<1$. Charges are extracted from the star and flow along such field lines with low energies. By contrast, if $\\alph>1$ or $\\alph<0$, a high voltage is generated, leading to unsteady $e^\\pm$ discharge on a scale-height smaller than the size of the polar cap. The discharge can power observed pulsars. Voltage fluctuations in the discharge imply unsteady twisting of the open flux tube and generation of Alfv\\'en waves. These waves are ducted along the tube and converted to electromagnetic waves, providing a new mechanism for pulsar radiation. ", "introduction": "Corotation of a plasma magnetosphere is impossible beyond the light cylinder of a star, and magnetic field lines that extend beyond this cylinder are twisted, $\\nabla\\times\\bB\\neq 0$. Thus, currents $\\bj_B=(c/4\\pi)\\nabla\\times\\bB$ are induced in the open magnetic flux tubes that connect the star (its ``polar caps'') to the light cylinder (Sturrock 1971). These currents are approximately force-free and flow along the magnetic field $\\bB$. A basic question of pulsar theory is what voltage develops along the open tube to maintain these currents. It determines the dissipated power and $e^\\pm$ creation that feeds the observed activity of pulsars. The customary pulsar model assumes that the electric current $\\jB$ extracted from the polar cap nearly matches $c\\rhoGJ$, where $\\rhoGJ=-{\\bf \\Omega}\\cdot\\bB/2\\pi c$ is the corotation charge density (Goldreich \\& Julian 1969). The deviation of current from $c\\rhoGJ$ was calculated as an eigen value of an electrostatic problem and found to be small (Arons \\& Scharlemann 1979). This is in conflict with recent global models of pulsar magnetospheres, which report $|\\jB-c\\rhoGJ|\\sim\\jB$ (e.g. Contopoulos et al. 1999; Spitkovsky 2006; Timokhin 2006; see Arons 2008 for a review). A significant mismatch between $\\jB$ and $c\\rhoGJ$ can be expected on general grounds (Kennel et al. 1979). Currents $\\jB$ are determined by the magnetic-field twisting near the light cylinder, while $\\rhoGJ$ is a local quantity at the polar cap that is practically independent of $\\jB$. The open tube is surrounded by the grounded closed magnetosphere\\footnote{ The closed magnetosphere with $\\jB=0$ is expected to have $\\rho=\\rhoGJ$ and $\\Epar\\approx 0$ (e.g. Arons 1979). } and may be thought of as a waveguide, filled with magnetized (1D) plasma. Compared to usual plasma-filled waveguides, it has two special features: (1) Current $\\jB$ is imposed on the tube. The twisted tube extends into the star, which is a good conductor and maintains $j_B$. ${\\rm sign}(\\jB)=\\pm 1$ corresponds to $\\pm$ charge flowing outward along the magnetic field lines. (2) Vacuum has effective charge density $\\rho_0=-\\rhoGJ$ as Gauss law in the rotating frame reads $\\nabla\\cdot\\bE=4\\pi(\\rho-\\rhoGJ)$. The key dimensionless parameter (which can vary along and across the tube) is \\be \\alph=\\frac{\\jB}{c\\rhoGJ}. \\ee In this Letter, we discuss basic properties of the polar-cap accelerator with arbitrary $\\alph$. First, we discuss what happens without $e^\\pm$ creation: \\S~\\ref{sec:slab} studies how the current is extracted from the polar cap of a radius $\\rpc$ and flows at small heights $z\\ll\\rpc$ (this region is called ``slab zone'' below), and \\S~\\ref{sec:tube} discusses how the flow extends to the region $z>\\rpc$ (``thin-tube zone''). We argue that the accelerator is inefficient if $0<\\alph<1$. The value of $\\alph$ depends largely on the angle $\\chi$ between ${\\bf\\Omega}$ and $\\bB$ at the polar cap. For aligned dipole rotators ($\\chi=0$), magnetospheric models predict $\\alph<1$ everywhere on the polar cap and $\\alpha<0$ near its edge (see Fig.~4 and 5 in Timokhin 2006). For orthogonal rotator ($\\chi\\approx\\pi/2$), $|\\alph|\\gg 1$ throughout most of the polar cap. Generally, the polar cap has three regions where $\\alph>1$, $0<\\alph<1$, and $\\alph<0$. We propose that pulsar activity originates in the polar-cap regions where $\\alph^{-1}<1$ (i.e. $\\alph>1$ or $\\alph<0$) as a high voltage is generated in these regions. We emphasize that the voltage is generated because $\\nabla\\times\\bB\\neq 0$, not because $\\rhoGJ\\neq 0$. The accelerator works as well if $\\rhoGJ=0$ (i.e. if $\\bf\\Omega\\perp\\bB$ at the polar cap), which corresponds to $\\alph\\rightarrow \\pm\\infty$. The accelerator height $h\\simlt\\rpc$ is regulated by unsteady $e^\\pm$ discharges (\\S~\\ref{sec:discharge}). \\S~\\ref{sec:Alfven} describes a mechanism for radio emission from unsteady discharges. ", "conclusions": "" }, "0710/0710.5734_arXiv.txt": { "abstract": "Big Bang nucleosynthesis requires a fine balance between equations of state for photons and relativistic fermions. Several corrections to equation of state parameters arise from classical and quantum physics, which are derived here from a canonical perspective. In particular, loop quantum gravity allows one to compute quantum gravity corrections for Maxwell and Dirac fields. Although the classical actions are very different, quantum corrections to the equation of state are remarkably similar. To lowest order, these corrections take the form of an overall expansion-dependent multiplicative factor in the total density. We use these results, along with the predictions of Big Bang nucleosynthesis, to place bounds on these corrections and especially the patch size of discrete quantum gravity states. ", "introduction": "\\label{sec:INTRODUCTION} Much of cosmology is well-described by a space-time near a spatially isotropic Friedmann--Robertson--Walker models with line elements \\begin{equation} \\md s^2= -\\md\\tau^2+a(\\tau)^2\\left(\\frac{\\md r^2}{1-kr^2}+r^2(\\md\\vartheta^2+ \\sin^2\\vartheta\\md\\varphi^2)\\right) \\end{equation} where $k=0$ or $\\pm 1$, sourced by perfect fluids with equations of state $P=w\\rho$. Such an equation of state relates the matter pressure $P$ to its energy density $\\rho$ and captures the thermodynamical properties in a form relevant for isotropic space-times in general relativity. Often, one can assume the equation of state parameter $w$ to be constant during successive phases of the universe evolution, with sharp jumps between different phases such as $w=-1$ during inflation, followed by $w=\\frac{1}{3}$ during radiation domination and $w=1$ during matter domination. Observationally relevant details can depend on the precise values of $w$ at a given stage, in particular if one uses an effective value describing a mixture of different matter components. For instance, during big bang nucleosynthesis one is in a radiation dominated phase mainly described by photons and relativistic fermions. Photons, according to Maxwell theory, have an exact equation of state parameter $w=\\frac{1}{3}$ as a consequence of conformal invariance of the equations of motion (such that the stress-energy tensor is trace-free). For fermions the general equation of state is more complicated and non-linear, but can in relativistic regimes be approximately given by the same value $w=\\frac{1}{3}$ as for photons. In contrast to the case of Maxwell theory, however, there is no strict symmetry such as conformal invariance which would prevent $w$ to take a different value. It is one of the main objectives of the present paper to discuss possible corrections to this value. For big bang nucleosynthesis, it turns out, the balance between fermions and photons is quite sensitive. In fact, different values for the equation of state parameters might even be preferred phenomenologically \\cite{FermionBoson}. One possible reason for different equations of state could be different coupling constants of bosons and fermions to gravity, for which currently no underlying mechanism is known. In this paper we will explore the possibility whether quantum gravitational corrections to the equations of state can produce sufficiently different values for the equation of state parameters. In fact, since the fields are governed by different actions, one generally expects different, though small, correction terms which can be of significance in a delicate balance. Note that we are not discussing ordinary quantum corrections of quantum fields on a classical background. Those are expected to be similar for fermions and radiation in relativistic regimes. We rather deal with quantum gravity corrections in the coupling of the fields to the space-time metric, about which much less is known a priori. Thus, different proposals of quantum gravity may differ at this stage, providing possible tests. An approach where quantum gravitational corrections can be computed is loop quantum cosmology \\cite{LivRev}, which specializes loop quantum gravity \\cite{Rov,ALRev,ThomasRev} to cosmological regimes. In such a canonical quantization of gravity, equations of state must be computed from matter Hamiltonians rather than covariant stress-energy tensors. Quantum corrections to the underlying Hamiltonian then imply corrections in the equation of state. This program was carried out for the Maxwell Hamiltonian in \\cite{MaxwellEOS}, and is done here for Dirac fermions. There are several differences between the treatment of fermions and other fields, which from the gravitational point of view are mainly related to the fact that fermions, in a first order formulation, also couple to torsion and not just the curvature of space-time. After describing the classical derivation of equations of state as well as steps of a loop quantization and its correction terms, we use big bang nucleosynthesis constraints to see how sensitively we can bound quantum gravity parameters. We have aimed to make the paper nearly self-contained and included some of the technical details. Secs.~\\ref{sec:canonical Formulation} on the canonical formulation of fermions, \\ref{sec:MODIFICATIONS} on quantum corrections from loop quantum gravity and \\ref{sec:BBN} on the analysis of big bang nucleosynthesis can, however, be read largely independently of each other by readers only interested in some of the aspects covered here. We will start with general remarks on the physics underlying the problem. ", "conclusions": "Big bang nucleosynthesis is a highly relativistic regime which, to a good approximation, implies identical equations of state for fermions and photons. There are, however, corrections to the simple equation of state $w=\\frac{1}{3}$ for fermions even classically. One observation made here is that the interaction term derived in \\cite{FermionImmirzi} leads to such a correction and might be more constrained by nucleosynthesis than through standard particle experiments \\cite{FermionTorsion}. We have not analyzed this further here because more details of the behavior of the fermion current would be required. A second source of corrections arises from quantum gravity. Remarkably, while quantum gravity effects on an isotropic background do correct the equations of state, they do so equally for photons and relativistic fermions. Initially, this is not expected for both types of fields due to their very different actions. Thus, quantum gravity effects do not spoil the detailed balance required for the scenario to work and bounds from big bang nucleosynthesis obtained so far are not strong. But there are interesting limits for the primary parameter, the patch size of a quantum gravity state. It is dimensionally expected to be proportional to the Planck length $\\ell_{\\rm P}$ but could be larger. In fact, current bounds derived here already rule out a patch size of exactly the elementary allowed value in loop quantum gravity. With more precise estimates, these bounds can be improved further. We have made use of quantum gravity corrections in a form which does not distinguish fermions from radiation. Although the most natural implementation, quite unexpectedly, provides equal corrections as shown here, there are several possibilities for differences which suggest several further investigations. Small deviations in the equations of state and thus energy densities of fermions and radiation are possible. First, there are always quantization ambiguities, and so far we tacitly assumed that the same basic quantization choice is made for the Maxwell and Dirac Hamiltonians. Such ambiguity parameters can be explicitly included in specific formulas for correction functions; see e.g.\\ \\cite{Ambig,ICGC,QuantCorrPert}. Independent consistency conditions for the quantization may at some point require one to use different quantizations for both types of fields, resulting in different quantum corrections and different energy densities. Such conditions can be derived from an analysis of anomaly-freedom of the Maxwell field and fermions coupled to gravity, which is currently in progress. As shown here, if this is the case it will become testable in scenarios sensitive to the behavior of energy density such as big bang nucleosynthesis. Moreover, assuming the same quantization parameters leads to identical quantum corrections for photons and fermions only on isotropic backgrounds. Small-scale anisotropies have different effects on both types of fields and can thus also be probed through their implications on the equation of state. For this, it will be important to estimate more precisely the typical size of corrections, which is not easy since it requires details of the quantum state of geometry. The crucial ingredient is again the patch size of underlying lattice states. On the other hand, taking a phenomenological point of view allows one to estimate ranges for patch sizes which would leave one in agreement with big bang nucleosynthesis constraints. Interestingly, corrections studied here provide upper bounds to the patch size, and other corrections from quantum gravity are expected to result in lower bounds. A finite window thus results, which can be shrunk with future improvements in observations." }, "0710/0710.3732_arXiv.txt": { "abstract": "% Edge-on spiral galaxies offer a unique perspective on disks. One can accurately determine the height distribution of stars and ISM and the line-of-sight integration allows for the study of faint structures. The Spitzer IRAC camera is an ideal instrument to study both the ISM and stellar structure in nearby galaxies; two of its channels trace the old stellar disk with little extinction and the 8 micron channel is dominated by the smallest dust grains (Polycyclic Aromatic Hydrocarbons, PAHs). \\cite{Dalcanton04} probed the link between the appearance of dust lanes and the disk stability. In a sample of bulge-less disks they show how in massive disks the ISM collapses into the characteristic thin dust lane. Less massive disks are gravitationally stable and their dust morphology is fractured. The transition occurs at 120 km/s for bulgeless disks. Here we report on our results of our Spitzer/IRAC survey of nearby edge-on spirals and its first results on the NIR Tully-Fischer relation, and ISM disk stability. ", "introduction": "For 32 edge-on galaxies, spanning Hubble type and mass, we fit the edge-on infinite disk model by \\cite{vdKruit81a} on the IRAC mocaics; the stellar dominated 4.5 $\\mu$m and the PAH emission at 8 $\\mu$m, with the stellar contribution subtracted \\citep[][]{Pahre04a}. The disk's total luminosity is inferred from the fitted model: $L_{disk} = 2 \\pi h^2 \\mu_0$, with $h$ the scale-length and $L_0$ the face-on central surface brightness \\citep[][]{Kregel02}. Figure \\ref{f:tf} shows the inferred Tully-Fischer relation for these disks for the stars (4.5 $\\mu$m). Notably, the slope ($\\alpha$) is 3.5, similar to what \\cite{Meyer06a} found but contrary for the increasing trend of slope with redder filters. The effects of age and metallicity of the stellar population become independent and opposite effects on the color-M/L relation in NIR \\citep[See][Fig. 2d]{BelldeJong}. Hence, the shallower slope in the IRAC stellar channels, could be the metallicity effect starting to dominate. ", "conclusions": "" }, "0710/0710.4262_arXiv.txt": { "abstract": "The accuracy of position measurements on stellar targets with the future Space Interferometry Mission (SIM) will be limited not only by photon noise and by the properties of the instrument (design, stability, etc.) and the overall measurement program (observing strategy, reduction methods, etc.), but also by the presence of other ``confusing'' stars in the field of view (FOV). We use a simple ``phasor'' model as an aid to understanding the main effects of this ``confusion bias'' in single observations with SIM. This analytic model has been implemented numerically in a computer code and applied to a selection of typical SIM target fields drawn from some of the Key Projects already accepted for the Mission. We expect that less than 1\\% of all SIM targets will be vulnerable to confusion bias; we show that for the present SIM design, confusion may be a concern if the surface density of field stars exceeds 0.4 star/arcsec$^2$. We have developed a software tool as an aid to ascertaining the possible presence of confusion bias in single observations of any arbitrary field. Some \\textit{a priori} knowledge of the locations and spectral energy distributions of the few brightest stars in the FOV is helpful in establishing the possible presence of confusion bias, but the information is in general not likely to be available with sufficient accuracy to permit its removal. We discuss several ways of reducing the likelihood of confusion bias in crowded fields. Finally, several limitations of the present semi-analytic approach are reviewed, and their effects on the present results are estimated. The simple model presented here provides a good physical understanding of how confusion arises in a single SIM observation, and has sufficient precision to establish the likelihood of a bias in most cases. We close this paper with a list of suggestions for future work on this subject. ", "introduction": "\\label{intro} The Space Interferometry Mission (SIM\\footnote{also currently called SIM--PlanetQuest}) is being designed by NASA's Jet Propulsion Laboratory to provide a facility-class instrument for measuring the positions and proper motions of stars at optical wavelengths with micro-arc-second (\\muas) precision. This represents an improvement by several orders of magnitude over the precision of all existing astrometric instruments. For faint sources (V $\\gtrsim 15$ mag.), SIM will also be more than a factor 10 better than any other future planned space mission, and will therefore uniquely permit many new classes of problems to be addressed. Such problems include: the direct measurement (for the first time) of the masses of earth-like planets in orbit around nearby stars; determining the distances to stars by direct triangulation over the whole Galaxy and out to the Magellanic Clouds; measuring the transverse motions of galaxies in the Local Group; and, establishing the shape of the dark matter distribution in the Galaxy. A description of the instrument and its current science program is available at the JPL/SIM web site.\\footnote{See {\\it http://planetquest.jpl.nasa.gov/SIM/sim\\_index.cfm} for descriptions of the current set of Key Projects. Almost half of the available 5-year Mission time is still unallocated.} The mission is now at the end of the detailed design phase. After more than 15 years of development, all the major technical questions have been answered. New devices have been invented in order to provide metrology internal to the spacecraft at a level of a few tens of picometers, a fraction of the inter-atomic distance in a molecule of oxygen. The next major step is to begin construction of the instrument. SIM will be the second optical interferometer in space devoted to astrometry, following the Fine Guidance Sensors (FGS) on the Hubble Space Telescope (HST). However, SIM is a Michelson interferometer using separated collectors, quite different from the filled-aperture white-light shearing interferometer design of the FGS. SIM includes three long-baseline interferometers housed on a common truss, each formed by two $\\approx 0.3$ m apertures which compress their light beams and guide them through delay lines to beam combiners. During operation, two interferometers are used for precision guiding of the spacecraft, and the third views the target of interest. Data are then accumulated by tracking the target until enough photons have been recorded to achieve the particular science goal. The precision with which astrometry can be done on a specific stellar target with SIM will be limited by photon noise, the design and stability of the instrument, and by the data calibration and processing. We have some control over the instrument properties, which depend on the specific choices made when implementing the design in hardware. Also, the operation and calibration of the instrument can be optimized so as to maximize the achievable precision. However, there is another source of error which may be present, and which is largely out of our control; it does not reduce the ultimate \\textit{precision} with which a given \\textit{single measurement} can be made with SIM, but it may reduce the final \\textit{accuracy} of that measurement. This source of error arises because of the presence of other stars in the SIM field of view (FOV) which can ``confuse'' any single observation made on the target star. The light from these extraneous stars perturbs the measurement, so that the measured target position can differ from the true position. The difference is a \\textit{bias} which can reach a level of many times the single-measurement precision estimated from the instrument parameters alone. It is this \\textit{single-measurement confusion bias} which concerns us in this paper. It is important to emphasize that the final \\textit{accuracy} of the astrometric parameters (position, parallax, and proper motion) determined by SIM for any given target star will be a result of carrying out a complex program of several single measurements on that target, plus repeated measurements on many other stars for the determination of calibration and baseline orientation parameters \\citep{bod04,miltur03}. Since we are concerned only with possible bias in a \\textit{single measurement}, the details of the entire observing program are not directly relevant here; however, it also means that we can not quantify the consequences of confusion bias to the final accuracy with which the ``end-of-mission'' astrometric parameters can be obtained on any specific target. It is clear that the effects of single-measurement confusion bias will generally diminish as more observations are combined. But this also means that projects which involve only a few observations of a target (e.g.\\ ``targets of opportunity'', single parallax measurements on nearby bright stars, etc.) may have a greater susceptibility to confusion bias. Specific aspects of confusion in astrometric measurements with SIM have been considered by several authors in the recent past. \\citet{dalalgriest01} showed that the characteristic response of SIM's fixed-baseline interferometer as a function of wavenumber and delay can be used to refine a model of the distribution of confusing stars in the FOV. This model can then be used to correct the measured position of the target of interest, and in many cases where the level of confusion is not too great the astrometric accuracy can approach the measurement precision. Dalal and Griest successfully applied their method to models of confused fields in the LMC in which $\\approx 16$ faint stars are scattered over the FOV around the $\\approx 19$ magnitude target star. Photon noise is included in these models. These authors then go on to consider the additional complication if one of the stars in the FOV changes brightness owing to a micro-lensing event, and show that an extension of their fitting algorithm to include the precision photometry provided by SIM's detectors permits even this apparently-intractable case to be handled almost as well. However, their method fails when the angular separation between any pair of sources in the FOV (as projected on the interferometer baseline) corresponds to a delay difference of $\\lesssim 2$ coherence lengths for the full bandpass of the detection system. This is a projected separation of 25 milli-arcseconds (mas). Indeed, this is a \\textit{general limit} for SIM observations. Our approach is somewhat simpler than that of Dalal and Griest, and yields some improvement in the minimum angular separation which can be measured, but the basic limitation can not be overcome. We will compare our approach to theirs in more detail in a future paper. \\citet{rajbokall01} also considered a number of specific cases of confusion on SIM astrometry. These authors introduced a graphical analogy using phasors as an aid to understanding how errors in the target position arise from confusing sources in the FOV. Typical target fields were constructed on a simulated image with grid spacing of 5 mas, and the amplitude and phase of the fringe which would be measured with SIM for a given wavelength on that image field was computed with a Fourier transform. Diffraction effects at the edges of the (presumed $\\approx 1''$ square) SIM FOV were included in constructing the model image, and vector averaging of the individual (narrow) SIM wavelength channels was used to simulate the 1-dimensional apodization of the fringes over the FOV caused by the decreasing coherence of the fringes as the bandwidth increases. \\citet{rajbokall01} were particularly interested in modeling the effects of mispointing of the FOV in subsequent visits to the same target field when target proper motions were being measured; in that case the actual distribution of field stars changes as some disappear from one side of the FOV and others appear at the other side. They included photon noise, and also addressed the issue of how the size of the FOV defined by the field stop influenced the level of confusion. Their source field models were constructed to simulate SIM observations of the position and proper motion of target stars in M31, the LMC, and the Galactic bulge. They concluded that the confusion-induced errors in position can often be significant (several times the measurement precision) for faint target stars but the proper motion errors are likely to be small. The errors are, as expected, smaller for wider measurement bands. In one other confusion-related study, \\citet{takvellin05} considered the effect of circumstellar disks on the measurement of stellar wobble during observations aimed at detecting extra-solar planets. Their models showed that neither the motion of the disk mass center nor the contamination by disk light is a serious threat to detecting planets around pre-main sequence stars; the basic reason for the insensitivity of these observations to confusion from circumstellar disks is that interferometers tend to resolve such an extended source, reducing its influence on the astrometry of the parent star. The studies summarized above have shown that confusion poses limits to the accuracy of any single SIM measurement, and have therefore succeeded in raising our awareness of this problem. However, the detailed design of SIM has changed since those studies were done, and many of the changes will affect on the modeling results. The size of the collector and its central obscuration, the entrance aperture field stop defining the geometrical FOV, the transmission efficiency of the optics, the fringe disperser design (which defines the bandwidth and central wavelength of the spectral channels), and the QE and spectral response of the detector are now all much better defined. Furthermore, previous studies have focused on specific science programs which were already suspected to be pushing the capabilities of the instrument; they have not provided us with any general ``tools'' for understanding and recognizing confusion, or for dealing with it. Previous studies have also often taken a statistical approach which is less suitable for answering direct questions about specific fields, such as: is a SIM observation of this particular target embedded in that particular field of stars likely to be confused? And, can the observation be done in such a way so as to reduce the confusion bias? What \\textit{a priori} information about the target and the field would help? And, if the observation has already been taken, can we identify the effects of confusion in the data? These questions have provided the motivation for the work described in this paper. This paper is organized as follows. In the next section, we summarize the basic Michelson interferometer response as it applies to SIM. We then recall the phasor model introduced by \\citet{rajbokall01} and elaborate upon it as a tool for understanding the behavior of confusion in SIM astrometry. Using this analytic model, together with updated knowledge of SIM's instrument parameters, we have constructed a simulation code for evaluating the likelihood of confusion bias in any specific field; details are presented in Section~\\ref{simu}. In Section~\\ref{limit_values}, we present single measurement confusion bias as a function of magnitude difference and projected separation of an additional star present within the SIM FOV. In Section~\\ref{applications}, we apply this semi-analytic model to a number of target fields drawn from the Key Projects which have already been chosen for the initial SIM science program. From this experience we then consider how the single-measurement confusion bias might be reduced through the addition of other information. The most useful additions appear to be knowledge of the approximate locations and spectral energy distributions (SEDs) of the target and of the most troublesome confusing stars in the SIM FOV. Finally, in Section \\ref{limitations}, we describe the limitations of our current approach. These limitations are primarily related to the simplified model of the focal plane of SIM which we have used here. In particular, in this paper, we have not modeled the detailed mechanism by which the spectral dispersion is implemented,\\footnote{A thin prism disperser turns SIM into an objective prism spectrograph on the CCD detector.} nor have we considered the pixellation of the focal plane by the CCD detector. We have explored these points with the aid of a more elaborate model that includes these effects, and find that the biases estimated using this more elaborate model differ only by small amounts from those provided by the approach described here.\\footnote {However, consideration of this more sophisticated model does suggest additional ways to reduce any confusion bias.} We have therefore chosen to present the main issues relevant to SIM confusion with a minimum of complication, and leave the discussion of the more elaborate instrument model to a future paper. Binary stars will be an important class of targets for SIM, and are the topic of one of the major Key Projects. In these cases, the two stars are in a bound orbit and are physically close to each other. Typical binaries to be studied with SIM will have separations from about a few mas to 1000 mas and orbital periods from a few days to several years. Stars in crowded fields can occasionally mimic the effects of binaries if their projected separations become small for particular baseline orientations, but the effects of confusion from such ``apparent'' binaries can be reduced (or even eliminated) by rotating the interferometer baseline and repeating the observation. However, for ``real'' binaries, rotation of the baseline is an integral part of the measurement process. The goal of the binary observation is to obtain the characteristics of the orbit, and this is done by measuring the positions of the components for a number of baseline orientations. These targets are sufficiently specialized that we have removed them from the list of crowded-field problems treated in this paper. Such observations treat binaries as ``signal\", whereas here we treat them as ``noise\". A discussion of astrometry on binary stars with SIM is planned for a future publication. ", "conclusions": "\\label{sum} We have examined the bias that can occur in a single measurement with SIM owing to the presence of field stars within the FOV. In order to accomplish this task we have presented a model for the SIM interferometer, and a description of how SIM carries out a single measurement of the position of an isolated target star. The measured instrument response is then perturbed by adding a field star to the model FOV; the difference in the angles measured in the two cases is called the ``confusion bias''. The extremes of this bias are calculated for the specific (but common) case of a binary system in order to illustrate its main properties. A number of source models are then developed which resemble the fields to be studied by SIM in several of the Key Projects already selected for inclusion in the initial mission science program. An unconfused version of the source model consisting only of the main target star is used as a reference measurement, and the results compared with a measurement made on the fully-populated field. The difference is the confusion bias in a single SIM measurement. Observations are simulated at various orientations of the interferometer baseline, and variants of the full field are examined in order to understand the sensitivity of the bias to structural details in the field. The magnitude of the confusion bias is found to depend on a number of factors, some obvious, others perhaps less so: \\begin{itemize} \\item the relative brightnesses of the target and the field stars; \\item the shapes of the SEDs of the target and field stars; \\item the angular separation of the stars from the center of the FOV; \\item the angular separation of the field stars from the target star as projected on the interferometer baseline; and, \\item the baseline orientation. \\end{itemize} The largest contributions to the confusion bias in a crowded field come from a small number of stars having small projected angular separations from the target, but these stars may actually be located outside of the FOV. Field stars which are less than 4 mag fainter than the target and which have projected separations within 100 mas of the target are potentially the most troublesome. The results of this study provides the understanding and the tools required to examine the likelihood of confusion bias in any single measurement with SIM. Unfortunately, data on the field stars in any specific FOV (especially their positions) is not likely to be available with sufficient accuracy to actually remove this bias.\\footnote{There are some possible exceptions one could imagine, but we have not explored them further here.} Our study nevertheless suggests some strategies for recognizing the presence of confusion bias and for dealing with it, both in the observation planning stage and in the data reduction stage. These strategies might include the following: \\begin{itemize} \\item While dealing with crowded fields, avoid fields with star densities in excess of 0.4 stars per square arcsec. \\item If avoidance is impossible, evaluate the likelihood of confusion in the field by using the tools developed here. \\item If confusion is likely, try to reduce your sensitivity to it by planning the observing program so that data is taken at the least sensitive orientations of the interferometer baseline. \\item If too little is known about the specific field, plan to distribute the available observing time over a number of orientations of the interferometer baseline which differ by a few degrees from each other. Inconsistent values in the data set can then be rejected with motivation. \\item If confusion is suspected in a given set of observations for which no prior data exists, acquiring new imagery from e.g., speckle or adaptive optics imaging would be useful for building a model. \\end{itemize} There is one additional strategy suggested by our more accurate model of the SIM focal plane. The CCD detector in the focal plane of SIM's camera is planned to have pixels which are smaller than the diameter of the FOV. If the data in the individual pixels can be made available, it would be possible to choose e.g.\\ only the central pixel, thereby effectively reducing the FOV and possibly attenuating an offending field star. The penalty of fewer target photons could then be offset by a reduction in the level of confusion bias. This possibility will be discussed in more detail in a future paper \\citep{sriron07}. We wish to emphasize that the results of this paper refer to a bias which may be present in a single measurement of angular position with SIM. The determination of the full set of astrometric parameters (position, parallax, proper motion) on any SIM target will be done with a number of measurements, reducing the effects of any single anomalous point. Furthermore, the ultimate accuracy of the results depends on an extensive calibration program to determine the instrumental parameters, including the baseline length and orientations for each field observed." }, "0710/0710.4054_arXiv.txt": { "abstract": "Boxy/peanut bulges in disk galaxies have been associated to stellar bars. In this talk, we discuss the different properties of such bulges and their relation with the corresponding bar, using a very large sample of a few hundred numerical N-body simulations. We present and inter-compare various methods of measuring the boxy/peanut bulge properties, namely its strength, shape and possible asymmetry. Some of these methods can be applied to both simulations and observations. Our final goal is to get correlations that will allow us to obtain information on the boxy/peanut bulge for a galaxy viewed face-on as well as information on the bars of galaxies viewed edge-on. ", "introduction": "Simulations have shown that bars are not vertically thin morphological features, but have a considerable vertical extent and a vertical structure, known as the Boxy/Peanut bulges (hereafter B/P; Combes \\& Sanders 1981, Combes et al. 1990). Comparisons between observations and $N$-body simulations have established this direct connection firmer (Athanassoula 2005 and references therein). Furthermore, observations have shown that both bars and B/P bulges are quite predominant in disc galaxies and that the corresponding frequencies are in good agreement with the link between the two structures (L\\\"utticke, Dettmar \\& Pohlen 2000). We measure the peanut properties in a large sample of several hundred $N$-body simulations ran by one of us (EA) for different purposes. More information on these simulations and on their properties can be found in Athanassoula \\& Misiriotis (2002) and Athanassoula (2003, 2007). In particular, we seek correlations between the properties of the bar and the properties of the B/P bulge. ", "conclusions": "We presented several methods to calculate the strength of the bar and the strength, shape and asymmetry of the B/P bulge and found strong correlations between their results. The most important correlation relates the strength of the bar with the strength of the B/P bulge, the strongest bars having the strongest peanuts. We also find that the strength of the peanut depends on the number of buckling episodes it underwent, the strongest bars having undergone more buckling episodes (Fig. 4). Finally, we find a very interesting result about $C_{2,z}(R)$, i.e. about the shape of the radial density profiles along cuts perpendicular to the equatorial plane. For strong bars, having a strong peanut or X-shaped bulge, this profile is more flat-topped, while for weaker bars, with more boxy-like bulges, it is more peaked. All the results summarised here are discussed in length by Athanassoula \\& Martinez-Valpuesta (2007, in preparation). \\begin{figure} \\begin{center} \\includegraphics[scale=0.35]{Martinez-Valpuesta_fig4.eps} \\caption{Correlations between bar and peanut properties. Each symbol corresponds to one simulation. The type of symbol is related to the number of buckling events suffered by the bar during its evolution. {\\it Left panel}: Strength of the B/P bulge measured with our Fourier based method vs. the strength of the bar. {\\it Right panel}: Shape of the B/P bulge (i.e. shape of the radial density profiles along cuts perpendicular to the equatorial plane, measured by the minimum of the kurtosis) plotted as a function of bar strength. } \\end{center} \\end{figure}" }, "0710/0710.0441_arXiv.txt": { "abstract": "We study the non-thermal emissions in a solar flare occurring on 2003 May 29 by using RHESSI hard X-ray (HXR) and Nobeyama microwave observations. This flare shows several typical behaviors of the HXR and microwave emissions: time delay of microwave peaks relative to HXR peaks, loop-top microwave and footpoint HXR sources, and a harder electron energy distribution inferred from the microwave spectrum than from the HXR spectrum. In addition, we found that the time profile of the spectral index of the higher-energy ($\\gsim 100$ keV) HXRs is similar to that of the microwaves, and is delayed from that of the lower-energy ($\\lsim 100$ keV) HXRs. We interpret these observations in terms of an electron transport model called {\\TPP}. We numerically solved the spatially-homogeneous {\\FP} equation to determine electron evolution in energy and pitch-angle space. By comparing the behaviors of the HXR and microwave emissions predicted by the model with the observations, we discuss the pitch-angle distribution of the electrons injected into the flare site. We found that the observed spectral variations can qualitatively be explained if the injected electrons have a pitch-angle distribution concentrated perpendicular to the magnetic field lines rather than isotropic distribution. % ", "introduction": "\\label{sec1} Observations of hard X-rays (HXRs), microwaves, and occasionally gamma-rays in solar flares tell us that a significant amount of non-thermal particles are produced. Among them, HXR and microwave observations are believed to provide the most direct information on electrons. Because HXRs below $\\sim 100$ keV are emitted primarily by electrons with energy below several hundred keV via {\\bremss} radiation \\citep{1971SoPh...18..489B}, whereas microwaves above $\\sim 10$ GHz are emitted by electrons above several hundred keV via {\\gyros} \\citep{1969ApJ...158..753R,1999spro.proc..211B}, these two sources of emission give us information on electrons in two different energy ranges. Therefore, a comparative study by using both HXR and microwave observations is useful for discussing the physics of flare non-thermal electrons over a wide range of energies. Impulsive behavior is commonly seen in both HXR and microwave lightcurves \\citep{1974IAUS...57..105K}, but the two emissions do not necessarily behave identically. Temporally, higher-energy HXR and microwave emissions tend to be delayed from lower-energy HXRs \\citep[e.g.,][]{1978ApJ...223..620C,1983Natur.305..292N,1985ApJ...292..699B,1997ApJ...487..936A}. \\cite{1997ApJ...487..936A} statistically analyzed the low-pass filtered HXR lightcurves for 78 flares observed with the {\\it Compton Gamma Ray Observatory} ({\\it CGRO}) and find a systematic increase of time delay toward higher energy. They interpreted these time delays in terms of electron precipitation under Coulomb collisions. Spatially, microwave sources do not always coincide with HXR sources. HXRs are typically emitted at the footpoint regions of the flare loop \\citep{1994PhDT.......335S} whereas microwaves are emitted mainly at the loop-top region \\citep{2002ApJ...580L.185M}. \\cite{2002ApJ...580L.185M} suggested that only electrons with a pancake pitch-angle distribution concentrated transverse to the magnetic field lines can explain the observed loop-top microwave source. Spectrally, \\cite{2000ApJ...545.1116S} statistically studied the correlation of the HXR and microwave spectral indices for 57 peaks of the non-thermal emission in 27 flares. They found that the electron energy distribution inferred from the microwave spectrum is systematically harder than that inferred from the HXR spectrum, and suggested that the electron energy distribution becomes harder towards higher energy. There are three probable explanations for such spectra: (1) two (or more) different electron populations with distinct physical characteristics, (2) ``second-step acceleration'' \\citep[e.g.,][]{1976SoPh...49..343B}, and (3) ``{\\TPP} (TPP)'' \\citep[e.g.,][]{1976MNRAS.176...15M}. \\cite{1976MNRAS.176...15M} presented analytic solutions of the electron energy continuity equation under two conditions: {\\it strong} and {\\it weak diffusion limits} \\citep{1966JGR....71....1K}. In the strong diffusion limit, electrons injected into the loop undergo significant scattering and then are quickly isotropized during the loop transit. They can escape from the loop with a precipitation rate proportional to their velocity, $\\nu_{\\rm p} \\propto v$. In the weak diffusion limit, on the other hand, electrons are less scattered during the transit. When the loss cone distribution is formed, the pitch-angle diffusion time $\\tau_{\\rm d}$, which is longer than the transit time, controls the electron precipitation, yielding $\\nu_{\\rm p} \\propto 1/\\tau_{\\rm d}$. The precipitation rate and the evolution of electrons vary, depending on which condition applies. There have been many observations that can be explained in terms of the TPP model \\citep[e.g.,][]{2000ApJ...531.1109L}. \\cite{2002ApJ...576L..87Y} reported the Nobeyama Radioheliograph (NoRH; \\citealt{1994PROCIEEE...82..705}) observation of a flare occurring on 1999 August 28. The NoRH observation showed clear flare-loop structure and propagating features along the loop. They showed that the microwave spectrum in the optically-thin regime is hard (with spectral index $\\sim 1.5$) around the loop-top, and then becomes softer (spectral index $\\sim 3.5$) toward the footpoints. Their observation indicates that the higher-energy electrons are efficiently trapped within the loop, supporting the TPP model. For this event, however, no HXR observation was available for comparison with the microwave observation. \\cite{2000ApJ...545.1116S} pointed out that the discrepancy of the energy distribution between the HXR and microwave emitting electrons found in their study could be explained by the TPP model. In their study, there was no imaging observation to confirm their suggestion. If the HXR and microwave sources do not coincide spatially, the discrepancy of the energy distribution between the HXR and microwave emitting electrons can be explained by the different spatial distribution between the HXR and microwave emitting electrons as a result of the TPP model. Imaging as well as spectral data at both HXR and microwave wavelengths are essential to confirm the role of TPP on the parent electrons In this paper we analyze the non-thermal emissions of the 2003 May 29 flare by using the {\\it Reuven Ramaty High Energy Solar Spectroscopic Imager} \\citep[RHESSI;][]{2002SoPh..210....3L}, the Nobeyama Radio Polarimeters \\citep[NoRP,][and references theirin]{1985PASJ...37..163N} and NoRH. RHESSI has superior spectroscopic ability from $\\sim 3$ keV to $\\sim 17$ MeV, providing the HXR spectrum from $\\sim 3$ keV to $\\sim 300$ keV with a spectral resolution of $\\sim 1$ keV and arbitrary energy bands. In previous studies, the temporal evolution of the (HXR) spectrum has been considered in less detail, probably due to instrumental limitations. However, the temporally-resolved analysis of the spectrum is important because non-thermal emissions and thus non-thermal electrons are the most ``time-varying'' objects in solar flares. RHESSI enables us to analyze an accurate, temporally-resolved HXR spectrum above $\\sim 100$ keV. Because the HXRs above $\\sim 100$ keV are mainly emitted by electrons above $\\sim 200$ keV \\citep{1996ApJ...464..974A}, RHESSI's well-resolved spectral data below $\\sim 300$ keV provides us more accurate information on electrons from tens to hundreds of keV than before. Combining the RHESSI HXR and NoRH/NoRP microwave spectral data allows us to fully cover the electrons from tens to thousands of keV. For a physical interpretation of the observations, we use a numerical model of TPP which treats the pitch-angle diffusion more generally than the analytic solutions developed for the weak and strong diffusion limits. \\cite{2000ApJ...543..457L} performed a similar treatment of the electron transport to explain their microwave observation of a flare on 1993 June 3. We also predict the microwave and HXR emissions from the calculated electron distribution. Comparing these model results with the observations, we discuss electron injection and transport, and address how the pitch-angle distribution of the injected electrons affects the evolution of the trapped and precipitating electrons, and their resultant emissions. The paper proceeds as follows. In {\\S}~\\ref{sec2} we present a comparative study of the non-thermal emissions of a solar flare occurring on 2003 May 29, by using the RHESSI HXR and Nobeyama microwave observations. Temporally-resolved spectra of the HXRs and microwaves are analyzed in detail. We discuss energy-dependent delays of the time profiles of the spectral indices, which have not been discussed in previous studies. In {\\S}~\\ref{sec3} we present our treatment of the TPP model. We numerically solve the spatially-homogeneous {\\FP} equation \\citep{1990A&A...230..213M} with the Coulomb interaction \\citep[e.g.,][]{1981ApJ...251..781L} and a time-dependent injection. In {\\S}~\\ref{sec4} we describe the time evolution of the trapped and precipitating electron distribution and the predicted microwave and HXR emissions. The behavior of the HXR and microwave emissions predicted by the model are compared with the observations, allowing us to give some constraints on the properties of flare non-thermal electrons. In {\\S}~\\ref{sec5} we conclude our study. % ", "conclusions": "\\label{sec5} We presented the comparative study of the non-thermal emissions of the flare occurring on 2003 May 29 using the RHESSI HXR and Nobeyama microwave observations. Further, we considered the electron transport model, TPP, to explain the observations. The 2003 May 29 flare showed two non-thermal HXR sources at the footpoints and a microwave source at the loop-top, as observed with RHESSI and NoRH. We interpreted this in terms of the TPP model. We presented the time profiles of the spectral indices of the higher-energy HXRs $\\gammat{H}{obs}$ as well as the lower-energy HXRs $\\gammat{L}{obs}$ and microwaves $\\alphat{obs}$. The spectra of microwaves and HXRs imply that the microwave emitting electrons have a harder energy distribution than the HXR emitting ones. % We found that the time profile of $\\gammat{H}{obs}$ shows similarity with that of $\\alphat{obs}$ rather than with $\\gammat{L}{obs}$, and is delayed from that of $\\gammat{L}{obs}$. We numerically solved the spatially-homogeneous {\\FP} equation for the TPP model to describe the evolution of electrons. Precipitating electrons have a softer energy distribution than the trapped ones in the weak diffusion regime. Differences of the injection pitch-angle distribution especially affect the evolution of the precipitating electrons. We calculated the microwave and HXR emissions from the calculated trapped electron distribution and precipitation flux for comparison with the observations. The TPP model in the weak diffusion regime can yield a soft HXR spectrum and a hard microwave spectrum. The calculated difference of the spectral indices between the HXRs and microwaves, $\\sim 1.5$, is in agreement with the observations. We further found that a pancake pitch-angle distribution for the injected electrons rather than an isotropic distribution is more adequate to qualitatively explain the temporal variation of $\\gammat{H}{obs}$. By comparing the model calculation with the observation, we can constrain the pitch-angle distribution of the injected electrons, which is crucially important for understanding the electron acceleration mechanism in solar flares. Currently, we are improving our treatment of the TPP model to include the spatial inhomogeneity in the Fokker-Planck equation. Using this, a systematic investigation of the best parameter set to explain the observation is in progress, and will be reported in the future." }, "0710/0710.0993_arXiv.txt": { "abstract": "The Alpha Magnetic Spectrometer (AMS), whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to measure charged cosmic ray spectra with energies up to the TeV region and with high energy photon detection capability up to a few hundred GeV, using state-of-the art particle identification techniques. It is equipped with several subsystems, one of which is a proximity focusing Ring Imaging \\CK\\ (RICH) detector equipped with a dual radiator (aerogel+NaF), a lateral conical mirror and a detection plane made of 680 photomultipliers and light guides, enabling precise measurements of particle electric charge and velocity ($\\Delta \\beta / \\beta \\sim$ 10${}^{-3}$ and 10${}^{-4}$ for $Z=$~1 and $Z=$~10~$-$~20, respectively) at kinetic energies of a few GeV/nucleon. Combining velocity measurements with data on particle rigidity from the AMS-02 Tracker ($\\Delta R / R \\sim$ 2\\% for $R=$~1~$-$~10 GV) it is possible to obtain a reliable measurement for particle mass. One of the main topics of the AMS-02 physics program is the search for indirect signatures of dark matter. Experimental data indicate that dark, non-baryonic matter of unknown composition is much more abundant than baryonic matter, accounting for a large fraction of the energy content of the Universe. Apart from antideuterons produced in cosmic-ray propagation, the annihilation of dark matter will produce additional antideuteron fluxes. Detailed Monte Carlo simulations of AMS-02 have been used to evaluate the detector's performance for mass separation, a key issue for $\\bar{D}/\\bar{p}$ separation. Results of these studies are presented. ", "introduction": "The Alpha Magnetic Spectrometer (AMS)\\cite{bib:ams}, whose final version AMS-02 is to be installed on the International Space Station (ISS) for at least 3 years, is a detector designed to study the cosmic ray flux by direct detection of particles above the Earth's atmosphere using state-of-the-art particle identification techniques. AMS-02 is equipped with a superconducting magnet cooled by superfluid helium. The spectrometer is composed of several subdetectors: a Transition Radiation Detector (TRD), a Time-of-Flight (TOF) detector, a Silicon Tracker, Anticoincidence Counters (ACC), a Ring Imaging \\CK\\ (RICH) detector and an Electromagnetic Calorimeter (ECAL). Fig.~\\ref{amsdet} shows a schematic view of the full AMS-02 detector. A preliminary version of the detector, AMS-01, was successfully flown aboard the US space shuttle Discovery in June 1998. \\begin{figure}[htb] \\center \\mbox{\\epsfig{file=ams2.eps,width=0.48\\textwidth,clip=}} \\caption{Exploded view of the AMS-02 detector.\\label{amsdet}} \\end{figure} The main goals of the AMS-02 experiment are: \\begin{itemize} \\item A precise measurement of charged cosmic ray spectra in the rigidity region between \\mbox{$\\sim$ 0.5 GV} and \\mbox{$\\sim$ 2 TV}, and the detection of photons with energies up to a few hundred GeV; \\item A search for heavy antinuclei ($Z \\ge$ 2), which if discovered would signal the existence of cosmological antimatter; \\item A search for dark matter constituents by examining possible signatures of their presence in the cosmic ray spectrum. \\end{itemize} The long exposure time and large acceptance (0.5 m${}^2$sr) of AMS-02 will enable it to collect an unprecedented statistics of more than $10^{10}$ nuclei. \\begin{figure}[htb] \\center \\mbox{\\epsfig{file=ev702white.eps,width=0.48\\textwidth,clip=}} \\vspace{-0.3cm} \\caption{A simulated proton event as seen in the AMS-02 display.\\label{amsdisplay1}} \\end{figure} ", "conclusions": "AMS-02 will provide a major improvement on the current knowledge of cosmic rays. A total statistics of more than 10${}^{10}$ events will be collected during its operation. Detailed simulations have been performed to evaluate the detector's particle identification capabilities, in particular those of the RICH, which might be crucial for the identification of an antideuteron flux resulting from neutralino annihilation. Simulation results show that the separation of light isotopes is feasible. Using a set of simple cuts based on event data, relative mass resolutions of $\\sim$~2 \\% and rejection factors up to 10${}^4$ have been attained in D/p separation at energies of a few GeV/nucleon. \\newpage" }, "0710/0710.5578_arXiv.txt": { "abstract": "We investigate properties of iron fluorescent line arising as a result of illumination of a black hole accretion disc by an X-ray source located above the disc surface. We study in details the light-bending model of variability of the line, extending previous work on the subject. We indicate bending of photon trajectories to the equatorial plane, which is a distinct property of the Kerr metric, as the most feasible effect underlying reduced variability of the line observed in several objects. A model involving an X-ray source with a varying radial distance, located within a few central gravitational radii around a rapidly rotating black hole, close to the disc surface, may explain both the elongated red wing of the line profile and the complex variability pattern observed in MCG--6-30-15 by {\\it XMM-Newton}. We point out also that illumination by radiation which returns to the disc (following the previous reflection) contributes significantly to formation of the line profile in some cases. As a result of this effect, the line profile always has a pronounced blue peak (which is not observed in the deep minimum state in MCG--6-30-15), unless the reflecting material is absent within the innermost 2--3 gravitational radii. ", "introduction": "Broad iron lines observed from many black hole systems most likely originate from the innermost regions of an accretion disc and their profiles are shaped by gravitational redshift and Doppler shifts. Modelling of the lines observed in several objects requires strongly enhanced fluorescent emission from a few gravitational radii (e.g., Wilms et al.~2001, Fabian et al.~2002, Miller et al.~2002, Miller et al.~2004; see review in Reynolds \\& Nowak 2003), which in turn indicates that a primary source of hard X-ray emission must also be located close to the black hole. \u00a0Thus, both the primary and reflected emission should be subject to strong gravity effects. These effects are also tentatively considered as an explanation of the complex variability pattern characterising radiation reflected from disc, including the iron line, and the primary hard X-ray continuum emission. Namely, weak variability of the reflected component, uncorrelated with the variability of the primary emission, has been reported in a number of sources (e.g., Vaughan \\& Fabian 2004; see review in Fabian \\& Miniutti 2005). This is contrary to expectations from a simple geometrical model of a hard X-ray source located close to the reflecting disc, where a strict correlation between variations of the primary and reflected emission should be observed. Fabian \\& Vaughan (2003) \u00a0first argued that such a reduced variability may be explained by relativistic effects, in particular by light bending and focusing the primary emission towards the accretion disc. Qualitatively, variations of the reflected emission should be much weaker as changes of the height of the X-ray source cause variations of its observed luminosity at infinity, while changes of the flux received by the disc (enhanced by the gravitational focusing; e.g., Matt et al.\\ 1992; Martocchia \\& Matt 1996; Petrucci \\& Henri 1997) are much weaker. However, for a static primary source located on the symmetry axis, the illuminating radiation is focused into the innermost part of the disc and the reflected emission is subject to similar light bending as the primary. As a result, similar variability characterises the primary and reflected emission, at least for observers with low inclination angles. Miniutti et al.\\ (2003) and Miniutti \\& Fabian (2004) have developed further this model to include rotation of the primary source around the axis and the resulting beaming of its emission toward outer regions of the disc. Then, variations of the reflected emission are reduced because of the more extended region it is produced. Predictions of their model are found to be consistent with observations of black-hole systems, e.g.\\ by Miniutti et al.\\ (2003), Miniutti, Fabian \\& Miller (2004), Fabian et al.\\ (2004). However, a number of important effects were not systematically studied. In particular, a specific pattern of motion of a primary source is assumed (including both location relative to the symmetry axis and azimuthal motion) but it is not discussed how strongly the resulting properties depend on these assumptions. For example, corotation of the primary source with the disc is assumed by Miniutti \\& Fabian (2004), which is likely close to the disc surface, but at most approximate at high latitudes. As the azimuthal motion of the primary source relative to the disc may affect significantly the reflected emission (see e.g.\\ Reynolds \\& Fabian 1997), it is not clear whether predictions of this model are specific to the underlying assumptions or generic to models involving the light bending. In this paper we systematically analyse the light bending model, concentrating on strong gravity effects. We neglect some other effects which may contribute to the original variability problem, for example, ionization of the disc surface (Nayakshin \\& Kazanas 2002). We focus here on the iron K$\\alpha$ line; an analysis of reflected emission including the Compton reflected radiation will be presented in our next paper. \u00a0 We study in details a number of geometrical scenarios of the source location and motion. We point out certain inadequacies in the original computations of Miniutti \\& Fabian (2004), and how they influence their quantitative results. We find also a novel scenario in which reduction of the line variability follows directly from properties of photon transfer in the Kerr metric. Namely, we find that a source located close to the disc surface, with a varying radial distance from a Kerr black hole, gives rise to both an approximately constant illumination of the surrounding disc and a very strong variability of the primary emission observed at infinity. Some aspects of variability in similar models have been considered recently e.g.\\ by Czerny et al.\\ (2004) and Pech\\'acek et al.\\ (2005), however, these studies considered only emission emerging locally from the region under the source. On the other hand, we find that transfer of primary emission to more distant regions of the disc is crucial for variability effects in such a model. We concentrate on low inclination objects, with application to Seyfert 1 galaxies in mind. Obviously, the Doppler shifts are more pronounced at high inclinations, but observational studies of high inclination objects are less advanced, either because they are obscured (as Seyfert 2 galaxies), or because the studies on dynamical time scale are not possible (as in stellar mass black hole systems). Furthermore, we consider only emission averaged over at least an orbital period. Effects resulting from varying azimuthal location of an off-axis source, with respect to observer, are studied, e.g., in Ruszkowski (2000), Yu \\& Lu (2000) and Goyder \\& Lasenby (2004). In Section 3 we analyse various physical effects relevant to formation of relativistic line profiles and variability effects; in Section 4 we apply our results to a Seyfert 1 galaxy MCG--6-30-15 which is the best studied object with clear signatures of strong gravity effects. ", "conclusions": "\\subsection{Variability models} We have extended the model, formulated by Miniutti \\& Fabian (2004), relating reduced variability of the reflected emission to changing magnitude of relativistic effects as location of the primary X-ray source changes. We find that original computations of Miniutti \\& Fabian (2004) - for a model involving a vertically moving source - overestimate the reduction effect by assuming a value of the outer radius of the disc which is too small for the range of the source heights considered in that model. On the other hand, we find a significant reduction of the variability of reflected emission in a model with a rapidly rotating black hole and a source moving radially, low above the disc surface. The reduced variability occurs then for the innermost range of radial distances, $\\le 4R_{\\rm g}$. We find also that - only in this range of parameters - the GR effects give rise to a significant decline around 6.5 keV in rms spectra. Note that generation of strong X-ray emission at these distances ($\\le 4R_{\\rm g}$) is consistent with the condition of a high value of $a$, as - for a rapidly rotating black hole - most accretion power is dissipated within a few innermost $R_{\\rm g}$. \\subsection{The black hole spin} Determination of the value of black hole spin or, even more fundamentally, verification of effects predicted by the Kerr metric solution of GR equations - remains a major issue of black hole astrophysics. In this context, several effects are taken into account for X-ray spectroscopy. Strong redshift of photons forming the observed red wings, at $E<4$ keV, is considered as evidence of rapid rotation of a black hole (e.g., Brenneman \\& Reynolds 2006). However, similar redshift can be obtained for $a=0$ if emission from $r_{\\rm d}<6$ is taken into account (Reynolds \\& Begelman 1997). Then, the derived high value of $a$ relies on assumption of no neutral iron emission from within the radius of marginal stability. Response of the line to increase of primary emission has been suggested for future studies. In particular, Reynolds et al.\\ (1999) indicate a bump occurring in the line profile and progressing to lower energies, with proceeding time, as a feature characteristic for a rapidly rotating black hole. Again, a similar - redshifted and Shapiro delayed - bump should appear for a non-rotating black hole if fluorescence inside $r_{\\rm ms}$ was taken into account. A straightforward analysis of the space-time metric could be performed for systems observed close to edge-on, for which effects due to lensing by a black hole would be directly seen in the line profile (e.g., Zakharov \\& Repin 2003). However, as noted in Narayan \\& McClintock (2005), there seems to be a selection effect preventing such systems from being observed. Then, we point out that a (largely unambiguous) analysis of imprints of the space-time metric would be possible in profile of the line resulting from irradiation by a strong flare just above the disc surface. If such a flare occurred at $r_{\\rm s} \\la 4$, properties of the space-time related with the black hole rotation would result in $\\ga 10$ per cent in magnitude effects in the line profile. The effects related to the value of $a$ are rather subtle but in principle possible to establish observationally. A compact flare dominating total emission would be required to make such an analysis feasible and viability of such scenario is uncertain. Such flares are indeed occasionally observed (e.g., Ponti et al.\\ 2004). Interestingly, however, the Fe K$\\alpha$ line was actually very weak during the flare analysed by Ponti et al.\\ (2004), and a strong line appeared in the spectrum with significant time delay after the flare. Finally, we emphasise that the reduced variability of reflected component - basing on mechanism advocated in this paper - is itself a direct manifestation of the nature of the Kerr space-time. Note that another effect resulting from properties of the Kerr metric, namely anisotropic emission of hard X-rays generated close to a rotating black hole, is qualitatively consistent with inclination-angle dependence of intrinsic spectra of Seyfert galaxies (Nied\\'zwiecki 2005). In support for the tentative relation of these two effects to properties of the Kerr metric, note also that comparison of the total mass in black holes in the local universe with the total luminosity produced by active galactic nuclei indicates that most supermassive black holes should rotate rapidly, e.g.\\ Elvis, Risaliti \\& Zamorani (2002). \\subsection{Modelling the red wing of relativistic Fe lines} As discussed above, detection of photons with $E < 4$ keV is considered as the evidence of rapid rotation of the black hole. Such strongly redshifted photons were revealed in the Fe line profiles observed in MCG--6-30-15 and several other objects (e.g., Miller et al.\\ 2004, Miniutti et al.\\ 2004). When fitted by models assuming a power-law radial emissivity, the observed profiles require $q>4$ within $(6-10) R_{\\rm g}$. Considering physical scenarios for generation of such profiles, we find that they \\newline (i) can be produced as a result of illumination by a hard X-ray source located close to a black hole (preferably at $r_{\\rm s} =2$--3) and rather close to the disc surface ($h_{\\rm s} \\la 1$); \\newline (ii) cannot be explained by models involving a source at a height of several $R_{\\rm g}$, or higher, close to the axis; \\newline if they are to come from a disc, with small inclination, surrounding a rotating black hole. Regarding (ii) we find, in agreement with previous studies, that such location of the source yields a steep $\\epsilon_{\\rm Fe}$, with $q>3$, in the innermost part of the disc. However, this steep $\\epsilon_{\\rm Fe}$ occurs only within $r_{\\rm d} \\le 2$ (and due to gravitational blueshift rather than light bending). The majority of emission from this region of the disc is either captured or bent toward the disc plane; moreover, the observed photons are strongly redshifted. As a result, this emission gives only a minor contribution to the line profile and it is not relevant to the shape and strength of the observed red wings. This conclusion is not affected by effects related with azimuthal motion of the source, returning radiation or angular emission law of fluorescence; these effects influence mostly the strength of the blue peak. On the other hand, the red wing is primarily related with position of the X-ray source. \\subsection{Returning radiation, non-local irradiation} A further challenge for modelling certain line profiles results from illumination of the surrounding disc (i.e. beyond a few $R_{\\rm g}$) by returning radiation. We find that this effect may be strong for the range of parameters (small $r_{\\rm s}$ and $V$) previously not explored. A similarly strong enhancement of Compton reflection by returning radiation, for very small distances ($r_{\\rm s} \\la 2$) of primary source, was noted recently by Suebsewong et al.\\ (2006). Another effect, related to bending of photon trajectories to the disc plane, affects models relating the Fe radial profile to some physical processes. These models typically assume that the radial profile of Fe emission is the same as some physically motivated profile of energy dissipation. E.g., Reynolds et al.\\ (2004) make such assumption in their analysis of the {\\it XMM} observation of MCG--6-30-15, applying model of the torqued-disc emission (where emission results mostly from extraction of rotational energy of the black hole and thus is very centrally concentrated). Moreover, they assume that there is no dissipation, and thus no reprocessing, beyond rather small (several $R_{\\rm g}$) outer radius. We emphasise that even for X-rays generated very low above the disc surface, a significant illumination of more distant regions of the disc should occur. Both the returning radiation and the non-locality of irradiation result primarily in enhancement of the blue peak of the line. \\subsection{Azimuthal motion} Impact of strong gravity is usually studied under specific assumptions on the X-ray source motion. Typically, angular velocity of the source is related to that of the disc (e.g. Ruszkowski 2000; Miniutti \\& Fabian 2004), while Dabrowski \\& Lasenby (2001) assume a static primary source. We note that change of $V$ significantly affects flux and profile of the iron line through combination of SR and GR effects. Moreover, the Kerr metric terms yield a non-trivial dependence between $V$ and $\\Omega$ (equation (\\ref{v})). Then, the assumed parametrisation of motion may be crucial for derived properties, e.g.\\ in some variability models. Obviously, additional effects may result from relativistic vertical or radial motion of the source (e.g.\\ Yu \\& Lu 2001), which were neglected in this paper. \\subsection{Applicability of our results} We focused above on MCG--6-30-15, for which several observed properties may be explained by our model. Similar effects, including reduced variability of reflection, pronounced red wings or strongly reflection dominated spectra, have been revealed in some Narrow Line Seyfert 1 galaxies (e.g. Fabian et al.\\ 2004, Ponti et al.\\ 2006), as well as in stellar mass black-hole systems (e.g.\\ Miniutti et al.\\ 2004) and in galactic nuclei with much higher black hole masses (Fabian et al.\\ 2005). Note that our scenario implies that the X-ray luminosity is underestimated by an order of magnitude. in these objects. On the other hand, most black-hole systems do not show signatures of such extreme effects. Then, similar reduction of the X-ray flux does not necessarily characterise any low inclination system. Note, however that such property would be inevitable for rapidly rotating black holes, where major fraction of accretion power is dissipated at very small $r_{\\rm s}$, if this power is converted into hard X-rays in situ. As in virtually all previous studies of that subject, we analysed impact of GR effects for the simplest case of an isotropic point source of hard X-ray emission, which approach follows from our lack of understanding of the X-ray source nature. The derived properties result from transfer of radiation from source to disc and observer and from disc to observer. Additional (but smaller in magnitude) effects may occur in more realistic models. E.g., Comptonization close to a Kerr black hole gives rise to anisotropic emission, cf.~Nied\\'zwiecki (2005), therefore radiation with different spectral index may irradiate various parts of the disc, while the transfer effects affect only normalisation and cut-off energy of the primary emission. The simplified description of X-ray source considered in this paper approximates most closely scenario with magnetic flares above the disc surface. Then, our results may be directly applicable to a model with flares occurring randomly at various radial distances. Qualitatively, we may assess similar variability effects for continuous spacial distributions of the hard X-ray source, e.g.\\ an extended corona covering the disc surface or a small hot torus replacing the disc within a few innermost $R_{\\rm g}$. Varying size of such an extended, hot plasma, in the Kerr metric, should give rise reduced variability of reflected component. In particular, a decline of direct emission from a shrinking corona or torus would be observed, even if its intrinsic luminosity remained unchanged, while increasing fraction of its emission would be bent to the disc plane giving rise to strong reflection component. The major shortcoming in our study results from the neglect of ionization effects. The strongest ionization should occur below the source, especially in models with low height above the disc surface. In general, strong ionization of this region should suppress the redshifted Doppler horns formed in the red-wing. On the one hand, this would make studies of effects related to value of $a$ less feasible. On the other hand, such depletion of the variable contribution to the red wing would reduce variations of the line profile at various flux states. However, details of ionization structure, and of the related reduction of the variable contribution to the lien, depend on additional assumptions, in particular, on azimuthal distribution of flares. Obviously, a strong single flare would give rise to stronger ionization than many weak flares uniformly distributed at a given $r_{\\rm s}$. \\subsection{Summary} If attributed to strong-gravity effects, both the time-averaged line profile and the variability pattern observed in MCG--6-30-15 by {\\it XMM-Newton\\/} independently indicate that a primary hard X-ray source must be located very close ($\\la 4R_{\\rm g}$) to a black hole, i.e.\\ in the region where Kerr metric effects become crucial. Rapid rotation of the black hole is necessary (and vastly sufficient) to account for reduction of variability of reflected emission for such spacial location of the source. Bending to the equatorial plane, underlying this reduction effect, appears to be the most pronounced effect of the Kerr metric to be studied in the X-ray spectra of black-hole systems." }, "0710/0710.5052_arXiv.txt": { "abstract": "Magnetic fluctuations generated by a tangling of the mean magnetic field by velocity fluctuations are studied in a developed turbulent convection with large magnetic Reynolds numbers. We show that the energy of magnetic fluctuations depends on magnetic Reynolds number only when the mean magnetic field is smaller than $B_{\\rm eq} / 4 {\\rm Rm}^{1/4}$, where $B_{\\rm eq}$ is the equipartition mean magnetic field determined by the turbulent kinetic energy and ${\\rm Rm}$ is magnetic Reynolds number. Generation of magnetic fluctuations in a turbulent convection with a nonzero mean magnetic field results in a decrease of the total turbulent pressure and may cause formation of the large-scale inhomogeneous magnetic structures even in an originally uniform mean magnetic field. This effect is caused by a negative contribution of the turbulent convection to the effective mean Lorentz force. The inhomogeneous large-scale magnetic fields are formed due to the excitation of the large-scale instability. The energy for this instability is supplied by the small-scale turbulent convection. The discussed effects might be useful for understanding the origin of the solar nonuniform magnetic fields, e.g., sunspots. ", "introduction": "Magnetic fields in astrophysics are strongly nonuniform (see, e.g., \\cite{M78,P79,KR80,ZRS83,RSS88,RH04,O03,BS05}). Large-scale magnetic structures are observed in the form of sunspots, solar coronal magnetic loops, etc. There are different mechanisms for the formation of the large-scale magnetic structures, e.g., the magnetic buoyancy instability of stratified continuous magnetic field \\cite{P79,P66,G70,P82}, the magnetic flux expulsion \\cite{W66}, the topological magnetic pumping \\cite{DY74}, etc. Magnetic buoyancy applies in the literature for different situations (see \\cite{P82}). The first corresponds to the magnetic buoyancy instability of stratified continuous magnetic field (see, e.g., \\cite{P79,P66,G70,P82}), and magnetic flux tube concept is not used there. The magnetic buoyancy instability of stratified continuous magnetic field is excited when the scale of variations of the initial magnetic field is less than the density stratification length. On the other hand, buoyancy of discrete magnetic flux tubes has been discussed in a number of studies in solar physics and astrophysics (see, e.g., \\cite{P82,P55,S81,SB82,FS93,SC94}). This phenomenon is also related to the problem of the storage of magnetic fields in the overshoot layer near the bottom of the solar convective zone (see, e.g., \\cite{SW80,T01,TH04,B05}). A universal mechanism of the formation of the nonuniform distribution of magnetic flux is associated with a magnetic flux expulsion. In particular, the expulsion of magnetic flux from two-dimensional flows (a single vortex and a grid of vortices) was demonstrated in \\cite{W66}. In the context of solar and stellar convection, the topological asymmetry of stationary thermal convection plays very important role in the magnetic field dynamics. In particular, the topological magnetic pumping is caused by the topological asymmetry of the thermal convection \\cite{DY74}. The fluid rises at the centers of the convective cells and falls at their peripheries. The ascending fluid elements (contrary to the descending fluid elements) are disconnected from one another. This causes a topological magnetic pumping effect allowing downward transport of the mean horizontal magnetic field to the bottom of a cell but impeding its upward return \\cite{ZRS83,DY74,GP83}. Turbulence may form inhomogeneous large-scale magnetic fields due to turbulent diamagnetic and paramagnetic effects (see, e.g., \\cite{KR80,Z57,VK83,K91,KR92,RKR03}). Inhomogeneous velocity fluctuations lead to a transport of mean magnetic flux from regions with high intensity of the velocity fluctuations. Inhomogeneous magnetic fluctuations due to the small-scale dynamo cause turbulent paramagnetic velocity, i.e., the magnetic flux is pushed into regions with high intensity of the magnetic fluctuations. Another effects are the effective drift velocities of the mean magnetic field caused by inhomogeneities of the fluid density \\cite{K91,KR92} and pressure \\cite{KP93}. In a nonlinear stage of the magnetic field evolution, inhomogeneities of the mean magnetic field contribute to the diamagnetic or paramagnetic drift velocities depending on the level of magnetic fluctuations due to the small-scale dynamo and level of the mean magnetic field \\cite{RK04}. The diamagnetic velocity causes a drift of the magnetic field components from the regions with a high intensity of the mean magnetic field. The nonlinear drift velocities of the mean magnetic field in a turbulent convection have been determined in \\cite{RK06}. This study demonstrates that the nonlinear drift velocities are caused by the three kinds of the inhomogeneities, i.e., inhomogeneous turbulence; the nonuniform fluid density and the nonuniform turbulent heat flux. The nonlinear drift velocities of the mean magnetic field cause the small-scale magnetic buoyancy and magnetic pumping effects in the turbulent convection. These phenomena are different from the large-scale magnetic buoyancy and magnetic pumping effects which are due to the effect of the mean magnetic field on the large-scale density stratified fluid flow. The small-scale magnetic buoyancy and magnetic pumping can be stronger than these large-scale effects when the mean magnetic field is smaller than the equipartition field determined by the turbulent kinetic energy \\cite{RK06}. The pumping of magnetic flux in three-dimensional compressible magnetoconvection has been studied in direct numerical simulations in \\cite{OS02} by calculating the turbulent diamagnetic and paramagnetic velocities. Turbulence may affect also the Lorentz force of the large-scale magnetic field (see \\cite{KRR89,KRR90,KR94,KMR96}). This effect can also form inhomogeneous magnetic structures. In this study a theoretical approach proposed in \\cite{KRR89,KRR90,KR94,KMR96} for a nonconvective turbulence is further developed and applied to investigate the modification of the large-scale magnetic force by turbulent convection and to elucidate a mechanism of formation of inhomogeneous magnetic structures. This paper is organized as follows. In Sect.~II we discuss the physics of the effect of turbulence on the large-scale Lorentz force. In Sect.~III we formulate the governing equations, the assumptions, the procedure of the derivations of the large-scale effective magnetic force in turbulent convection. In Sect.~IV we study magnetic fluctuations and determine the modification of the large-scale effective Lorentz force by the turbulent convection. In Sect.~V we discuss formation of the large-scale magnetic inhomogeneous structures in the turbulent convection due to excitation of the large-scale instability. Finally, we draw conclusions in Sect.~VI. In Appendix~A we perform the derivation of the large-scale effective Lorentz force in the turbulent convection. ", "conclusions": "In the present study we investigate magnetic fluctuations generated by a tangling of the mean magnetic field in a developed turbulent convection. When the mean magnetic field $B \\ll B_{\\rm eq} / 4 {\\rm Rm}^{1/4}$, the energy of magnetic fluctuations depends on magnetic Reynolds number. We study the modification of the large-scale magnetic force by turbulent convection. We show that the generation of magnetic fluctuations in a turbulent convection results in a decrease of the total turbulent pressure and may cause formation of the large-scale magnetic structures even in an originally uniform mean magnetic field. This phenomenon is due to a negative contribution of the turbulent convection to the effective mean magnetic force. The large-scale instability causes the formation of inhomogeneous magnetic structures. The energy for these processes is supplied by the small-scale turbulent convection, and this effect can develop even in an initially uniform magnetic field. In contrast, the Parker's magnetic buoyancy instability is excited when the density stratification scale is larger than the characteristic scale of the mean magnetic field variations (see \\cite{P66,P79,G70}). The free energy in the Parker's magnetic buoyancy instability is drawn from the gravitational field. The characteristic time of the large-scale instability is of the order of the Alfv\\'{e}n time based on the large-scale magnetic field. We study an initial stage of formation of the large-scale magnetic structures for horizontal and vertical mean magnetic fields relative to the vertical direction of the gravity field. In the turbulent convection there are two ranges for the large-scale instability of the horizontal mean magnetic field. The first range for the instability is related to the negative contribution of turbulence to the effective magnetic pressure for the case of $L_{\\rho} < L_{B}$, while the second range for the instability is mainly caused by the anisotropic contribution of the turbulent convection to the effective magnetic force. The large-scale instability of the vertical uniform mean magnetic field is caused by the modification of the mean magnetic tension by small-scale turbulent convection. The discussed effects in the present study might be useful for the understanding of the origin of the sunspot formation. Since in the present study we neglect very small Brunt-V\\\"{a}is\\\"{a}l\\\"{a} frequency based on the gradient of the mean entropy, we do not investigate the large-scale dynamics of the mean entropy. This problem was addressed in \\cite{KM00} whereby the modification of the mean magnetic force by the turbulent convection was not taken into account. In order to study magnetic fluctuations and the modification of the large-scale Lorentz force by turbulent convection we apply the spectral $\\tau$ approximation (see Sect.~III). The $\\tau$ approach is an universal tool in turbulent transport that allows to obtain closed results and compare them with the results of laboratory experiments, observations and numerical simulations. The $\\tau$ approximation reproduces many well-known phenomena found by other methods in turbulent transport of particles and magnetic fields, in turbulent convection and stably stratified turbulent flows (see below). In turbulent transport, the $\\tau$ approximation yields correct formulae for turbulent diffusion, turbulent thermal diffusion and turbulent barodiffusion (see, e.g., \\cite{EKR96,BF03}). The phenomenon of turbulent thermal diffusion (a nondiffusive streaming of particles in the direction of the mean heat flux), has been predicted using the stochastic calculus (the path integral approach) and the $\\tau$ approximation. This phenomenon has been already detected in laboratory experiments in oscillating grids turbulence \\cite{EEKR04} and in a multi-fan turbulence generator \\cite{EEKR06} in stably and unstably stratified fluid flows. The experimental results obtained in \\cite{EEKR04,EEKR04} are in a good agreement with the theoretical studies performed by means of different approaches (see \\cite{EKR96,PM02}). The $\\tau$ approximation reproduces the well-known $k^{-7/3}$-spectrum of anisotropic velocity fluctuations in a sheared turbulence (see \\cite{EKRZ02}). This spectrum was found previously in analytical, numerical, laboratory studies and was observed in the atmospheric turbulence (see, e.g., \\cite{L67}). In the turbulent boundary layer problems, the $\\tau$-approximation yields correct expressions for turbulent viscosity, turbulent thermal conductivity and the classical heat flux. This approach also describes the counter wind heat flux and the Deardorff's heat flux in convective boundary layers (see \\cite{EKRZ02}). These phenomena have been studied previously using different approaches (see, e.g., \\cite{MY75,Mc90,Z91}). The theory of turbulent convection \\cite{EKRZ02} based on the $\\tau$-approximation explains the recently discovered hysteresis phenomenon in laboratory turbulent convection \\cite{EEKRM06}. The results obtained using the $\\tau$-approximation allow also to explain the most pronounced features of typical semi-organized coherent structures observed in the atmospheric convective boundary layers (\"cloud cells\" and \"cloud streets\") \\cite{ET85}. The theory \\cite{EKRZ02} based on the $\\tau$-approximation predicts realistic values of the following parameters: the aspect ratios of structures, the ratios of the minimum size of the semi-organized structures to the maximum scale of turbulent motions and the characteristic lifetime of the semi-organized structures. The theory \\cite{EKRZ02} also predicts excitation of convective-shear waves propagating perpendicular to the convective rolls (\"cloud streets\"). This waves have been observed in the atmospheric convective boundary layers with cloud streets \\cite{ET85}. A theory \\cite{ZEKR07} for stably stratified atmospheric turbulent flows based on the $\\tau$-approximation and the budget equations for the key second moments, turbulent kinetic and potential energies and vertical turbulent fluxes of momentum and buoyancy, is in a good agrement with data from atmospheric and laboratory experiments, direct numerical simulations and large-eddy simulations (see detailed comparison in Sect. 5 of \\cite{ZEKR07}). The detailed verification of the $\\tau$ approximation in the direct numerical simulations of turbulent transport of passive scalar has been recently performed in \\cite{BK04}. In particular, the results on turbulent transport of passive scalar obtained using direct numerical simulations of homogeneous isotropic turbulence have been compared with that obtained using a closure model based on the $\\tau$ approximation. The numerical and analytical results are in a good agreement. In magnetohydrodynamics, the $\\tau$ approximation reproduces many well-known phenomena found by different methods, e.g., the $\\tau$ approximation yields correct formulae for the $\\alpha$-effect \\cite{KR80,RK93,RK00,RKR03}, the turbulent diamagnetic and paramagnetic velocities \\cite{Z57,VK83,K91,KR92,RKR03}, the turbulent magnetic diffusion \\cite{KR80,VK83,KRP94,RKR03,RK04}, the ${\\bf \\Omega} {\\bf \\times} {\\bf J}$ effect and the $\\kappa$-effect \\cite{KR80,RKR03}, the shear-current effect \\cite{RK03,RK04,RK07}. Generation of the large-scale magnetic field in a nonhelical turbulence with an imposed mean velocity shear has been recently investigated in \\cite{BH05} using direct numerical simulations. The results of these numerical simulations are in a good agreement with the theoretical predictions based on the $\\tau$ approximation (see \\cite{RK03,RK04,RK07}) and with the numerical solutions of the nonlinear dynamo equations performed in \\cite{BS05B,RKL06} (see detailed comparison in \\cite{RK07}). The validity of the $\\tau$ approximation has been tested in the context of dynamo theory, in direct numerical simulations in \\cite{BSM05}. The alpha effect in mean field dynamo theory becomes proportional to a relaxation time scale multiplied by the difference between kinetic and current helicities. It is shown in \\cite{BSM05} that the value of the relaxation time is positive and, in units of the turnover time at the forcing wavenumber, it is of the order of unity. Kinetic and current helicities are shown in \\cite{BSM05} to be dominated by large scale properties of the flow. Recent studies in \\cite{SSB07} of the nonlinear alpha effect showed that in the limit of small magnetic and hydrodynamic Reynolds numbers, both the second order correlation approximation (or first-order smoothing approximation) and the $\\tau$ approximation give identical results. This is also supported by simulations \\cite{BS07} of isotropically forced helical turbulence whereby the contributions to kinetic and magnetic alpha effects are computed. The study performed in \\cite{BS07} provides an extra piece of evidence that the $\\tau$ approximation is a useable formalism for describing simulation data and for predicting the behavior in situations that are not yet accessible to direct numerical simulations." }, "0710/0710.0663_arXiv.txt": { "abstract": "The intrinsic anisotropy $\\delta$ and flattening $\\epsilon$ of simulated merger remnants is compared with elliptical galaxies that have been observed by the SAURON collaboration, and that were analysed using axisymmetric Schwarzschild models. Collisionless binary mergers of stellar disks and disk mergers with an additional isothermal gas component, neglecting star formation cannot reproduce the observed trend $\\delta = 0.55 \\epsilon$ (SAURON relationship). An excellent fit of the SAURON relationship for flattened ellipticals with $\\epsilon \\geq 0.25$ is however found for merger simulations of disks with gas fractions $\\geq 20\\% $, including star formation and stellar energy feedback. Massive black hole feedback does not strongly affect this result. Subsequent dry merging of merger remnants however does not generate the slowly-rotating SAURON ellipticals which are characterized by low ellipticities $\\epsilon < 0.25$ and low anisotropies. This indicates that at least some ellipticals on the red galaxy sequence did not form by binary mergers of disks or early-type galaxies. We show that stellar spheroids resulting from multiple, hierarchical mergers of star-bursting subunits in a cosmological context are in excellent agreement with the low ellipticities and anisotropies of the slowly rotating SAURON ellipticals and their observed trend of $\\delta$ with $\\epsilon$. The numerical simulations indicate that the SAURON relation might be a result of strong violent relaxation and phase mixing of multiple, kinematically cold stellar subunits with the angular momentum of the system determining its location on the relation. ", "introduction": "A popular formation scenario for early-type galaxies is the collision and merger of two roughly equal-mass galaxies with mass ratios between 1:1 and 4:1. This famous major merger scenario \\citep{1972ApJ...178..623T} has been very successful in explaining observed properties of ellipticals, like their kinematics, surface density profile or isophotal shape \\citep{1981MNRAS.197..179G,1983MNRAS.205.1009N,1988ApJ...331..699B, 1990Natur.344..379B,1992ARA&A..30..705B,1992ApJ...400..460H, 1993ApJ...409..548H,1993A&A...278...23B,1999ApJ...523L.133N,2001ApJ...554..291C, 2003ApJ...597..893N,2005MNRAS.357..753G,2005MNRAS.360.1185J, 2005A&A...437...69B,2006MNRAS.369..625N,2006MNRAS.372..839N, 2006ApJ...641...21R,2006astro.ph..7446C}. Numerical simulations for example showed that the family of disky, fast rotating ellipticals could result from stellar disk galaxy mergers with unequal mass ratios of 3:1 to 4:1 \\citep{1998giis.conf..275B,1998ApJ...502L.133B,1999ApJ...523L.133N, 2003ApJ...597..893N,2005A&A...437...69B,2006MNRAS.372..839N} or from gas-rich 1:1 to 2:1 disk mergers where the gas subsequently settles into the equatorial plane of the merger remnant and produces a new stellar disk component \\citep{2005MNRAS.359.1379K, 1996ApJ...471..115B,2001ApJ...555L..91N,2002MNRAS.333..481B, 2005ApJ...622L...9S}. Boxy, slowly rotating ellipticals, on the other hand form in stellar disk-disk mergers with mass ratios of 1:1 to 2:1 \\citep{1994ApJ...427..165H,2003ApJ...597..893N} or from multiple major disk mergers \\citep{1996ApJ...460..101W}. \\citet{2007arXiv0709.3439B} showed that repeated minor mergers result in remnant properties very similar to one corresponding to major mergers. A serious problem of the major merger scenario was the fact that cosmological models do not predict a dependence of the mass ratio of mergers on total galaxy mass or luminosity \\citep{2003ApJ...597L.117K}. If, on the other hand, the mass ratio determines the isophotal shape and rotational properties of merger remnants one would expect that the ratio of the number of fast rotating, disky ellipticals to the number of slowly rotating, boxy systems should be independent of luminosity \\citep{2006ApJ...636L..81N}. This is in contrast with observations that show a strong dependence of isophotal shape and rotational properties on galaxy mass. While massive galaxies are preferentially boxy, slow rotators, lower-mass ellipticals are predominantly disky and fast rotators (for a summary see \\citealp{1996ApJ...464L.119K}). \\citet{2003ApJ...597L.117K} argued that this mass dependence of galaxy properties could be explained as a result of differences in the morphologies of the merging progenitors (see e.g. \\citealp{2007MNRAS.tmp..808K}). Using semi-analytical models they showed that gas-rich disk-disk mergers dominate at the low-mass end of ellipticals while intermediate mass ellipticals should have formed preferentially from mixed mergers involving a disk and an elliptical galaxy. Finally, the most massive early-type galaxies should have experienced a last elliptical-elliptical merger (dry merger) \\citep{2006ApJ...636L..81N}. Mixed mergers have not yet been studied in details, although, according to \\citet{2003ApJ...597L.117K} they should be more frequent than dry mergers (see however \\citealp{2007arXiv0706.1243H}). Dry mergers and their implications for the formation of the red galaxy sequence have however received a lot of attention recently \\citep{2005MNRAS.359.1379K,2005MNRAS.361.1043G,2005astro.ph..6044F, 2006ApJ...636L..81N,2006MNRAS.369.1081B}. Further refinement of theoretical models has recently been achieved by including black-hole physics in simulations of galaxy mergers and elliptical galaxy evolution. Energetic feedback from central black holes might solve some pending problems of major merger models, like the suppression of late inflow of cold gas and star formation that would make ellipticals look much bluer than observed \\citep{2005ApJ...620L..79S}. In summary, despite several still unsolved questions (e.g. \\citealp{2007astro.ph..2535N}), the major merger scenario has become a popular model in order to explain the origin of bulge-dominated, spheroidal galaxies (see e.g. \\citealp{2007arXiv0706.1246H,2007arXiv0706.1243H}). Progress in understanding galactic evolution is often driven by strong interactions between observers and theorists. Increasingly more sophisticated theoretical/numerical models are confronted with continuously improving observations that lead to new theoretical challenges. One example is the SAURON project \\citep{2004MNRAS.352..721E} which aims to determine the 2-dimensional structural and kinematical properties of early-type galaxies using a panoramic integral-field spectrograph. In order to interpret the observations and study the intrinsic galaxy structure, axisymmetric Schwarzschild models are applied to the observations \\citep{2006MNRAS.366.1126C,2007MNRAS.379..418C}. The results, published so far, have revealed interesting fine structures and physical properties which provide new and deeper insight into the origin of galaxies, in particular when compared to simulation \\citep{2000MNRAS.316..315B,2007MNRAS.376..997J}. In this paper we confront a recently published SAURON analysis of preferentially axisymmetric elliptical galaxies \\citep{2007MNRAS.379..418C} with the predictions of numerical merger simulations and cosmological models of galaxy formation. Section 2 summarizes the observations. Section 3 shows that simulations of collisionless and gaseous disk-disk mergers, neglecting star formation and stellar feedback cannot reproduce the observational results. We demonstrate that star formation and stellar energy feedback has a strong effect on the final structure of merger remnants, leading to a good agreement with the SAURON observations of fast rotating ellipticals. The origin of the round, almost isotropic, slowly rotating SAURON ellipticals is explored in section 4. Isolated, dry mergers of ellipticals that formed as discussed in section 3 cannot explain these objects. We show however that cosmological initial conditions, leading to a series of multiple major and minor mergers, coupled with local star bursts generate spheroidal stellar systems in very good agreement with the observations. Conclusions follow in section 5. ", "conclusions": "The numerical simulations, discussed in the previous sections, have shown that interstellar gas dynamics, star formation and stellar feedback plays a crucial role in order to reproduce the observed kinematical and isophotal properties of fast rotating, early-type galaxies. The final structure of the merger remnants depends on the initial mass ratio and gas fraction. The remnants are more round, less anisotropic and more rotationally supported the smaller the mass ratio $M_1/M_2 \\geq 1$ of the progenitors and the larger the initial gas fraction. The dependence of $\\delta$ on $\\epsilon_{int}$ is in agreement with the observed trend found in the SAURON sample. Subsequent dry re-merging of disk-disk merger remnants however does not generate the observed slowly-rotating red SAURON ellipticals with small anisotropies and ellipticities. This indicates that at least some early-type galaxies on the red galaxy sequence formed in a different way. We find that multiple mergers of stellar substructures that formed from cold gas infall into dark matter halos in cosmological simulations produce round, isotropic and slowly-rotating relaxed stellar systems that are in perfect agreement with the SAURON observations. Multiple mergers of stellar systems in dense group enviroments therefore appear to be a promising alternative scenario for the origin of the red, massive galaxy population. Despite the fact that merger simulations with star formation lead to a correlation between anisotropy and ellipticity ($\\delta = 0.67 \\times \\epsilon_{int}$) that is very similar to that inferred from observations its origin is not understood yet. It is interesting that merger remnants appear to move closer towards this relation along lines of constant $V/\\sigma$ (i.e. roughly constant specific angular momentum) in the case of a strong relaxation process. Here, strong relaxation is defined as the merger of a system of kinematically cold systems of stars that lateron break up and generate a kinematically hot stellar remnant. Several conditions could lead to such a violent dynamical process. The cold stellar clumps could for example have formed in the star-bursting tidal tails of interacting, gas-rich disk galaxies. Another possibility is the cosmological multiple merging of dark matter substructures with an embedded stellar systems. The SAURON relation might represent the relaxed and phase-mixed end state of these complex mergers with the location of the remnant on the relation being determined by its specific angular momentum which is related to its value of $V/\\sigma$. More theoretical work will be required in order to better understand these interesting questions and their connection to early-type galaxy formation." }, "0710/0710.2450_arXiv.txt": { "abstract": "Smooth double crossing of the phantom barrier $w_{\\Lambda} = -1$ has been found possible in cosmological model with Gauss-Bonnet-scalar interaction, in the presence of background cold dark matter. Such crossing has been observed to be a sufficiently late time phenomena and independent of the sign of Gauss-Bonnet-scalar interaction. The luminosity distance versus redshift curve shows a perfect fit with the $\\Lambda CDM$ model up to $z=3.5$. ", "introduction": "The puzzle associated with recent cosmic acceleration, triggered by $70\\%$ of dark energy or more \\cite{a1} is far from being resolved uniquely. In the mean time, cosmologists are being confronted with yet another more intriguing challenge to explain the crossing of the so called phantom divide line $(w_{\\Lambda}=-1)$, at sufficiently late time of cosmological evolution. Some recent analysis \\cite{a1},\\cite{b1} of the presently available observational data are in favour of the value $w_{de}<-1$, at present., $w_{de}$ being the dark energy equation of state. There are also a lot of evidence all around \\cite{c}, of a dynamical dark energy equation of state, which has crossed the so called phantom divide line $w_{\\Lambda}=-1$ recently, at the value of red-shift parameter $z \\approx 0.2$. Apparently though the problem turns out to be more serious and complicated, but then, the puzzle of crossing the phantom divide line has also rendered some sort of selection rule. $\\Lambda CDM$-model, which is known to suffer from the disease of fine tuning (see \\cite{d} for a comprehensive review) can now be ruled out due to the requirement of a dynamic state parameter. Further, if the analysis of Vikman \\cite{e} is correct, then it is not possible to cross the phantom divide line in a single minimally coupled scalar field theory, without violating the stability both at the classical \\cite{f} and also at the quantum mechanical levels \\cite{g}, (though it has recently been inferred \\cite{on} that quantum Effects which induce the $w<-1$ phase, are stable in the $\\phi^4$ model). Thus single minimally coupled scalar field models like quintessence $(w>-1)$ and phantom $(w<-1)$ are to be kept aside. Consequently, we are now left with some what more complicated models. One of these is a hybrid model, composed of two scalar fields, viz, quintessence and phantom - usually dubbed as quintom model \\cite{h}. Other models like non-minimal scalar tensor theory of gravity \\cite{i}, hessence \\cite{j} and models including higher order curvature invariant terms \\cite{k} also exist in the literature. \\\\ Gauss-Bonnet term is yet another candidate which may be pursued for the purpose. The possibility of crossing the phantom divide line through Gauss-Bonnet interaction has been explored in some recent works \\cite{l},\\cite{m}. But then, these models are even complicated in the sense that either brane-world scenario \\cite{l} or scalar field and matter coupling \\cite{m} are invoked. In this article the possibility of smooth crossing of the phantom divide line $w_{\\Lambda}=-1$ has been expatiated simply by introducing Gauss-Bonnet-Scalar coupling term in the 4-dimensional Einstein-Hilbert action.\\\\ Gauss-Bonnet term arises naturally as the leading order of the $\\alpha'$ expansion of heterotic superstring theory, where, $\\alpha'$ is the inverse string tension \\cite{n}. Gauss-Bonnet term is topologically invariant and thus does not contribute to the field equations in four dimensions. However, the low energy limit of the string theory gives rise to the dilatonic scalar field which is found to be coupled with various curvature invariant terms \\cite{o}. The leading quadratic correction gives rise to Gauss-Bonnet term with a dilatonic coupling \\cite{p}. Therefore it is reasonable to consider Gauss-Bonnet interaction in four dimension with dilatonic-scalar coupling. Several works with Gauss-Bonnet-dilatonic coupling are already present in the literature \\cite{q}. In particular, important issues like - late time dominance of dark energy after a scaling matter era and thus alleviating the coincidence problem, crossing the phantom divide line and compatibility with the observed spectrum of cosmic background radiation have also been addressed recently \\cite{km}.\\\\ In a recent work with Gauss-Bonnet interaction \\cite{a}, a solution in the form $a=a_{0}e^{A\\sqrt t}$ ($a$ being the scale factor, and $A>0$) has is been found to satisfy the field equations with different forms (sum of exponentials, sum of inverse exponentials, sum of powers and even quadratic) of potentials. Solution in a more general form $(a=a_{0}e^{A t^f}), A>0, 00, 00$. Then our strategy is to explore wide regions for the values of the other four parameters, by means of a bi-dimensional analysis in the $m_0 - m_{1/2}$ plane for different fixed values of $A_0$ and tan$\\beta$. It is important to mention that we are not presenting an exhaustive search in all the possible regions, but we concentrated on those regions which have received more attention in the literature, see for example \\cite{belanger}. In Fig.~(\\ref{tb10new}), we present the results for tan$\\beta=10$, for three values of $A_0$, namely $A_0=1000, 0, -1000$ GeV, shown in the top, middle and bottom panels respectiveley. The yellow region (lower right corner) is where the $\\tilde{\\tau}$ is the LSP, the lighter and darker areas (red and blue for the online version in colours) define the allowed regions for the EC and AC respectively according to the observed DM density. The area of the EC region depends on the size of the interval of values of the parameter $\\alpha$, the lower and upper bounds of $\\alpha$ determine the upper and lower boundaries of the EC region. As can be seen from the figure, the region where both criteria are fullfilled is very small, in fact, only for the highest values of $\\alpha$ there is an intersection between both criteria. This behavior holds for all values of $A_0$ in the interval $[-1000, 1000]$ GeV, here we are showing only the extreme and central values. \\begin{figure}\\centering \\includegraphics[height=15cm, width=4.8cm, viewport=200 0 300 350]{mytb_10.pdf} \\caption{Allowed regions in the parameter space for AC (lighter gray/red) and EC (darker grey/blue) for the mSUGRA model with sgn$\\mu=+$, tan$\\beta=10$, and $A_0=1000$ GeV, top panel, $A_0=0$ GeV, middle panel, and $A_0=-1000$ GeV, bottom panel. The figures show the so called bulk and coannihilation regions. The yellow region shows where the stau is the LSP.} \\label{tb10new} \\end{figure} Repeating the same procedure for larger values of tan$\\beta$, it is found that the intersection region for both criteria becomes larger, but it gets to be significant for the largest values of this parameter. This is clearly shown in Fig.~(\\ref{tb50new}), which is equivalent to Fig.~(\\ref{tb10new}), but for tan$\\beta=50$. In this case the bottom panel is for $A_0=-500$ GeV. It is clear from the figure that for these values of tan$\\beta$ both criteria are consistent, as shown by the large intersection area for values of $A_0$ in the interval $[0,1000]$ GeV. For negative values of $A_0$ the intersection region decreases with $A_0$, see the bottom panel of the figure. For even lower values of $A_0$ the intersection becomes insignificant. \\begin{figure}\\centering \\includegraphics[height=15cm, width=4.8cm, viewport=200 0 300 350]{mytb_50.pdf} \\caption{The same as Fig.~(\\ref{tb10new}), but for tan$\\beta=50$, and now $A_0=-500$ GeV in the bottom panel.} \\label{tb50new} \\end{figure} In Figs.~(\\ref{results2}) we present the same analysis but for the Focus Point region, and for the central value $A_0=0$. The situation is consistent with the previous results, both criteria intersect for $\\tan \\beta=50$ and there is nearly no intersection for $\\tan \\beta=10$. \\begin{figure}\\centering \\includegraphics[width=8cm]{fig3.pdf} \\includegraphics[width=8cm]{fig4.pdf} \\caption{Allowed regions in the parameter space for AC (lighter gray/red) and EC (darker grey/blue) in the mSUGRA model with $A_0=0$, sgn$\\mu=+$, tan$\\beta=10$, top panel, and tan$\\beta=50$, bottom panel. The region shows the so called Focus Point region.} \\label{results2} \\end{figure} This analysis allows us to arrive to one of the main results of our work. The use of both criteria favours large values of tan$\\beta$. In Figs.(\\ref{fig:Mhiggsvsm12}) and (\\ref{fig:chi-higgs1}) we show the allowed values for the LSP and the Higgs mass after constraining the parameter space with the abundance and entropy criteria. As can be seen from Fig. (\\ref{fig:Mhiggsvsm12}), the current limit for the Higgs favour, combined with the AC and EC criteria, favours even more a large value of $\\tan \\beta$. This, in turn, puts a constraint on the allowed SUSY mass spectra of the bulk and coannihilation regions: it gives an LSP of mass $m_{\\chi} \\sim 140$ GeV for $\\tb ~ 10$, and a lower bound for the LSP mass $m_{\\chi} \\gtrsim 150$ GeV for large $\\tb$. \\begin{figure}\\centering \\includegraphics[width=8cm]{mhiggsvsm12.pdf} \\caption{Allowed values for $M_{Higgs}$ as function of $m_{1/2}$. As can be seen from the figure, the present bound on the Higgs mass favours a large $\\tan \\beta$.} \\label{fig:Mhiggsvsm12} \\end{figure} \\begin{figure}\\centering \\centerline{\\includegraphics[width=8cm,angle=0]{fig6.pdf}} \\caption{The figure shows the LSP mass plotted versus the Higgs mass, points above the dashed line are the allowed values for the LSP. The points in blue correspond to the Focus Point region, the ones in red to the bulk and coannihilation regions.} \\label{fig:chi-higgs1} \\end{figure} Further analysis, which is currently under way, is required to give more precise conclusions about this new method to constrain the parameter space of the mSUGRA model \\cite{nuevo}. \\vspace{0.3cm} We acknowledge partial support by CONACyT\\\\ M\\'exico, under grants 32138-E, 34407-E and 42026-F, and PAPIIT-UNAM IN-122002, IN117803 and \\\\ IN115207 grants. JZ acknowledges support from DGEP-UNAM and CONACyT scholarships." }, "0710/0710.0619_arXiv.txt": { "abstract": "The large majority of EGRET point sources remain without an identified low-energy counterpart, and a large fraction of these sources are most likely extragalactic. Whatever the nature of the extragalactic EGRET unidentified sources, faint unresolved objects of the same class must have a contribution to the diffuse extragalactic gamma-ray background (EGRB). Understanding this component of the EGRB, along with other guaranteed contributions from known sources, is essential if we are to use this emission to constrain exotic high-energy physics. Here, we follow an empirical approach to estimate whether a potential contribution of unidentified sources to the EGRB is likely to be important, and we find that it is. Additionally, we show how upcoming GLAST observations of EGRET unidentified sources, as well as of their fainter counterparts, can be combined with GLAST observations of the Galactic and extragalactic diffuse backgrounds to shed light on the nature of the EGRET unidentified sources even without any positional association of such sources with low-energy counterparts. ", "introduction": "\\label{intro} The origin of the isotropic diffuse emission (Sreekumar et al.\\ 1998) in energies between $100 {\\, \\rm MeV}$ and $20 {\\, \\rm GeV}$, detected by the {\\it Energetic Gamma-Ray Experiment Telescope} (EGRET) aboard the {\\it Compton Gamma-Ray Observatory}, remains one of the great unknowns of GeV-energy astrophysics. There are two major questions that still remain unanswered. 1.\\ How much of the diffuse emission detected by EGRET is, in fact, extragalactic, and what is the spectrum of this extragalactic background? And 2.\\ what fraction of the extragalactic emission can be attributed to each of the observationally established classes of gamma-ray emitters? Despite the associated uncertainties, these two issues are critical in any attempt to use gamma-ray observations to constrain exotic high-energy physics and yet-undetected classes of theorized gamma-ray emitters. To answer the first question, a good understanding of the Galactic diffuse emission in the EGRET energy range is required. In order to obtain the intensity and spectrum of the extragalactic emission from the EGRET sky maps, the Galactic emission needs to be modeled and subtracted. This is made complicated by the discrepancy between the observed Milky Way spectrum in energies of $\\gtrsim 1 {\\rm GeV}$ and theoretical expectations (Hunter et al.\\ 1997). The observed spectrum is more shallow than model predictions based on the local demodulated cosmic ray spectrum. This deviation is known as the ``GeV excess'', and although various explanations have been proposed to account for part or all of the discrepancy, its origin remains a matter of debate (e.g., Pohl et al.\\ 1997; B\\\"{u}sching et al.\\ 2001; Strong et al.\\ 2004b; Kamae et al.\\ 2005; de Boer et al.\\ 2006; Strong 2006; Stecker \\etal 2007). As a result, determinations of the gamma-ray background using different Galactic emission models yield very different answers, both in intensity and in spectrum, despite being based on the same set of observations (e.g., Sreekumar \\etal 1998; Strong et al.\\ 2004a). Attempts to answer the second question have been plagued by uncertainties in the cosmic density and evolution of the two established classes of extragalactic gamma-ray emitters: normal galaxies and blazars. Our observational knowledge of the gamma-ray properties of normal galaxies is very limited, as the sample of normal galaxies which have been observed in gamma rays consists of only the Milky Way and a marginal detection of the Large Magellanic Cloud (Sreekumar et al.\\ 1992; Hunter et al.\\ 1997; Hartman et al.\\ 1999). For this reason, the accuracy of theoretical estimates of the contribution of normal galaxies to the gamma-ray background is unavoidably at the order-of-magnitude level (e.g., Lichti et al.\\ 1978; Pavlidou \\& Fields 2002). But even in the case of blazars, which are by far the most numerous and best studied class of identified gamma-ray emitters, estimates of their contribution to the gamma-ray background vary from a few percent to $100\\%$ of the background originally reported by the EGRET team (e.g., Padovani \\etal 1993; Stecker \\& Salamon 1996a; Kazanas \\& Perlman 1997; Mukherjee \\& Chiang 1999; M\\\"{ucke} \\& Pohl 2000; Narumoto \\& Totani 2006; Dermer 2007). The issue is further complicated by the existence of 171 sources which, at the time of publication of the 3rd EGRET catalog (hereafter 3EG; Hartman \\etal 1999), had not been positively or potentially associated with a lower-energy counterpart. These sources are collectively known as the unidentified EGRET sources, and they are more numerous than any established group of gamma-ray emitters. The distribution of these sources on the sky is such that a Galactic feature can be clearly distinguished - however a large number of sources are located away from the Galactic plane and the Galactic center\\footnote{Note however that the presence of sources at high latitudes does not, in itself, constitute proof that these objects are extragalactic (see e.g.\\ the Gould Belt discussion in Gehrels et al.\\ 2000).}. No more than a handful of sources can be associated with the Milky Way halo if the Milky Way is not many times brighter in gamma-rays than similar galaxies such as M31 (Siegal-Gaskins \\etal 2007). Hence, it is almost certain that the EGRET unidentified sources include a significant extragalactic component. Although the nature of these extragalactic sources remains unknown, it is reasonable to believe that there is a large number of fainter, unresolved objects of the same class, which are guaranteed to have {\\em some} contribution to the extragalactic gamma-ray background (EGRB). If these sources represent yet unidentified members of some known class of gamma-ray emitters (e.g.\\ blazars), then excluding them from any calculation of the contribution of the parent class to the diffuse background would lead to a significantly underestimated result due to an incorrect normalization of the bright-end of the gamma-ray luminosity function. If they represent an unknown class of gamma-ray emitters, then the contribution of their unresolved counterparts to the diffuse emission would significantly limit the diffuse flux left to be attributed to known classes, exotic processes, and truly diffuse emission. Hence, some contribution of unresolved unidentified sources to the EGRB is certain. It is therefore clear that until we either answer the question of the nature of unidentified sources or derive some strong constraint indicating that a possible contribution of such unresolved objects to the EGRB would indeed be minor, we cannot hope to fully understand the origin of the EGRB. Detailed predictions for the level of the unidentified source contribution to the EGRB involve important uncertainties: since no low-energy counterparts have been identified, we have no estimates of distance, and therefore no estimates of the gamma-ray luminosities of these sources. As a result, very few constraints can be placed on their cosmic distribution and evolution. However, very simple estimates can offer some guidance on whether ignoring this EGRB component may be a safe assumption to make. For example, we can use the number of unidentified sources, the minimum flux resolvable by EGRET, and the observed intensity to the extragalactic gamma-ray background to place rough limits on the distance scales associated with resolved and unresolved unidentified sources so that unresolved sources do not overproduce the background. A population of unbeamed, non-evolving, single-luminosity sources uniformly distributed in Euclidian space are resolvable out to a distance $D$ by an instrument of number flux sensitivity $F_{\\rm min}$. The relation between $D$, $F_{\\rm min}$, and number luminosity $L$ in this case is simply $L=4\\pi D^2F_{\\rm min}$. If the instrument detects $N$ such sources, their number density $n_{\\rm source}$ can be estimated to be $n_{\\rm source} = 3N/4\\pi D^3$. If the same distribution of sources continues out to a distance $d>D$, the isotropic intensity (photons per unit area per unit time per unit solid angle) from the {\\em unresolved} members of this population will be $I_{\\rm unres} = \\int_D^d dI_{\\rm shell}$, where $dI_{\\rm shell} = (1/4\\pi) (n_{\\rm source}4\\pi r^2 dr) L/(4\\pi r^2)$ is the contribution from sources within a spherical shell located at a distance $r$ from the observer. Substituting our results for $n_{\\rm source}$ and $L$ above, and performing the integral, we obtain \\begin{equation} I_{\\rm unres} = 3NF_{\\rm min}(d-D)/4\\pi D\\,. \\end{equation} If we require that the unresolved emission from this population does not exceed the EGRB observed by EGRET ($I_{\\rm unres} \\leq I_{\\rm EGRB}$), we obtain $I_{\\rm EGRB} \\geq 3NF_{\\rm min}(d-D)/4\\pi D$. Substituting $N\\sim 100$ for the population of extragalactic unidentified sources (see discussion in \\S \\ref{samples}), $F_{\\rm min} \\sim 10^{-7} {\\rm \\, ph \\, cm^{2} \\, s^{-1}}$ for the sensitivity of EGRET, and $I_{\\rm EGRB} \\sim 10^{-5} {\\rm \\, ph \\, cm^{2} \\, s^{-1} \\, sr^{-1}}$ (Sreekumar et al. 1998), we obtain $d \\lesssim 6D$. This result implies that the largest distance out to which such a distribution of objects persists cannot be larger than a few times the distance out to which these objects are currently resolved, since in any other case these sources would overproduce the EGRB. It is therefore conceivable that the unidentified source contribution to the EGRB is significant, if not dominant. In this work, we approach the problem from a purely empirical point of view. Instead of attempting to {\\em predict} the level of a diffuse component due to unresolved objects of the same class as unidentified EGRET sources, we try to assess whether there are any empirical indications that this component is, in fact, minor. We construct samples of unidentified sources which, based on their sky distribution, are likely to consist mostly of extragalactic objects. Under the assumption that the majority of these sources can be treated as members of a single class of gamma-ray emitters, we seek to answer the following three questions: (1) Is it likely that unresolved objects of the same class could have a significant contribution to the EGRB at least in some energy range? (2) How would the collective spectrum of their emission compare to the measured spectrum of the EGRB deduced from EGRET observations? (3) How are GLAST observations expected to improve our understanding of the nature of unidentified sources, based on the insight gained from our analysis? This paper is structured as follows. In \\S \\ref{samples} we discuss the samples of resolved unidentified sources used in our analysis. Our formalism for constructing the collective emission spectrum of unresolved unidentified sources is presented in \\S \\ref{formalism}. Inputs from EGRET data used in our analysis are described in \\S \\ref{inputs}. In \\S \\ref{results} we describe our results, and in \\S \\ref{sec:GLAST} we discuss how these results are expected to improve once GLAST observations become available. Finally, we summarize our conclusions in \\S \\ref{sec:disc}. ", "conclusions": "\\label{sec:disc} In this work, we have used a purely empirical model to explore the possibility that unresolved gamma-ray sources of the same class as unidentified EGRET sources have an appreciable contribution to the EGRB. We have argued that some unidentified source contribution to the gamma-ray background is guaranteed. We have additionally found that (1) if most high-latitude unidentified sources are assumed to be extragalactic, a one order of magnitude extension of the cumulative flux distribution to lower energies without breaks implies a significant contribution to the EGRB, at least at the lower part of the EGRET energy range; and (2) the spectrum of the cumulative emission of such unresolved sources would be very consistent with the observational determination of Strong et al.\\ (2004a) of the EGRET EGRB within systematics. We emphasize that the purpose of this study is not to estimate, even at the order-of-magintude level, the diffuse flux expected from extragalactic unresolved unidentified sources, but rather to place constraints on the flux distribution of these objects under the constraint that the measured background should not be exceeded in any energy interval. Our treatment is therefore different in purpose and spirit from most past work aiming to estimate the level of the contribution of different populations to the extragalactic gamma-ray background. Such work has been traditionally of two types. The first involves population models built from some understanding of the physics of the sources (such as, e.g., the models of M\\\"{u}cke \\& Pohl 2000 and Dermer 2007 in the case of blazars; Miniati 2003, Keshet et al.\\ 2003, Blasi, Gabici, \\& Brunetti 2007 in the case of clusters of galaxies; Lichti, Bignami \\& Paul 1978 and Pavlidou \\& Fields 2002 in the case of normal galaxies). The second involves models of the population luminosity function based on our knowledge of the source population from other wavelengths and normalized to fit EGRET data (such studies require a sample of detected, identified members, and are therefore applicable to blazars only, e.g. Chiang et al.\\ 1995, Stecker \\& Salamon 1996, Mukherjee \\& Chiang 1999, Narumoto \\& Totani 2006). In our case, lacking any knowledge of the physics of sources as well as of even the bright end of the luminosity function (since in absence of identifications no redshift and hence no luminosity can be derived for any of the sources), we have reversed the problem. Instead of using some assumed cosmic evolution for the sources to derive the expected level of contribution to the EGRB, as in all of the investigations mentioned above, we have used the tightest possible constraint on the allowable EGRB contribution of these sources (the observed EGRET background) to constrain the flux distribution of the unresolved sources. Our conclusions for the overall expected unresolved unidentified source intensity come from the observation that our constraints on the flux distribution are indeed very tight; hence, the contribution of the unidentified sources to the EGRB is likely to be high. Our analysis suggests that any model of the EGRB would be incomplete without some treatment of the unidentified source contribution. The results of our empirical model therefore motivate the pursuit of specific population models for the unidentified sources. Although such models involve a more restrictive set of assumptions and increased uncertainties, they can provide more concrete predictions for the luminosity function of unresolved objects. Once a luminosity function model is assumed, and under the assumption that unresolved members of this class do indeed contribute most of the extragalactic diffuse emission, additional estimates can be made regarding the distance scales associated with this population. The dynamical range in fluxes of the {\\em resolved} members of the population, as can be seen in Fig. \\ref{lognlogs}, is smaller than an order of magnitude. For a single-luminosity population, this corresponds to a factor of 3 dynamical range in distance, which increases or decreases if the typical source luminosity increases or decreases with increasing redshift respectively. Such arguments can constrain not only the distance scales, but also the number density, intrinsic brightness, and evolution of the resolved and unresolved objects, if in a model-dependent fashion. Additional constraints and predictions, once a luminosity function has been assumed, can be derived through the multi-messenger and multi-wavelength approach. If the emission from extragalactic unidentified sources is primarily of hadronic origin, it will be accompanied by neutrinos at comparable fluxes, and that may have potentially observable consequences in the TeV range for future km$^3$ neutrino detectors such as IceCube and Km3Net. If on the other hand the gamma-ray emission is primarily leptonic, it will be accompanied by X-ray emission, and their extragalactic background flux in the X-ray band may provide an additional constraint. We will pursue such models and calculations in an upcoming publication. In this work we have tried, where possible, to make assumptions, which, if anything, {\\em underestimate} the possible contribution of unresolved unidentified sources to the EGRB. An exception to this general trend is our working assumption that the majority of the resolved, high-latitude 3EG unidentified sources belong to a single class. It is conceivable that instead, the resolved unidentified sources are a collection of members of several known and unknown classes of gamma-ray emitters. In this case, it is still likely that the summed contribution of unresolved members of all parent classes to the diffuse background is significant. However, the construction of a single cumulative flux distribution from all sources and its extrapolation to lower fluxes is no longer an indicative test for the importance of such a contribution. Although we have argued that the contribution of unresolved unidentified sources to the EGRB is likely to be important or even dominant, the cumulative emission spectrum derived here and presented in Fig.\\ \\ref{ress} is only an upper limit, as it was derived demanding that the observed EGRB is not exceeded at any energy above 100 MeV. The contribution of unresolved unidentified sources is further constrained by allowing for the presence of the guaranteed contributions of unresolved normal galaxies and blazars. Finally, it is noteworthy that we have found no evidence for an inconsistency between the population properties of high-latitude unidentified sources and those of blazars. The presence of a yet-unknown population of extragalactic high-energy emitters among the high-latitude EGRET unidentified sources remains one of the most tantalizing possibilities in GeV astronomy and one of the most exciting prospects for the GLAST era. However, the presently available data on the spectral index distribution and the cumulative flux distribution of these sources (this work), as well as their variability properties (e.g., Nolan et al.\\ 2003), are consistent with their being members of the blazar population. If indeed a large number of members of the blazar class are present among the entragalactic unidentified sources, then this could have potentially serious effects on our understanding of the redshift distribution of resolved blazars and consequently of the blazar luminosity function." }, "0710/0710.0796_arXiv.txt": { "abstract": "We present the result of a photometric and Keck-LRIS spectroscopic study of dwarf galaxies in the core of the Perseus Cluster, down to a magnitude of M$_{\\rm B}$ $= -12.5$. Spectra were obtained for twenty-three dwarf-galaxy candidates, from which we measure radial velocities and stellar population characteristics from absorption line indices. From radial velocities obtained using these spectra we confirm twelve systems as cluster members, with the remaining eleven as non-members. Using these newly confirmed cluster members, we are able to extend the confirmed colour-magnitude relation for the Perseus Cluster down to M$_{\\rm B}$ $= -12.5$. We confirm an increase in the scatter about the colour magnitude relationship below M$_{\\rm B}$ $= -15.5$, but reject the hypothesis that very red dwarfs are cluster members. We measure the faint-end slope of the luminosity function between M$_{\\rm B}$ $= -18$ and M$_{\\rm B}$ $= -12.5$, finding $\\alpha$ $= -1.26$ $\\pm$ 0.06, which is similar to that of the field. This implies that an overabundance of dwarf galaxies does not exist in the core of the Perseus Cluster. By comparing metal and Balmer absorption line indices with $\\alpha$-enhanced single stellar population models, we derive ages and metallicities for these newly confirmed cluster members. We find two distinct dwarf elliptical populations: an old, metal poor population with ages $\\sim$ 8 Gyr and metallicities $[{\\rm Fe/H}]$ $<$ $-0.33$, and a young, metal rich population with ages $<$ 5 Gyr and metallicities $[{\\rm Fe/H}]$ $>$ $-0.33$. Dwarf galaxies in the Perseus Cluster are therefore not a simple homogeneous population, but rather exhibit a range in age and metallicity. ", "introduction": "Faint, low mass galaxies are the most numerous galaxy type in the universe, and are thus fundamental in understanding galaxy formation. However, their low luminosities and low surface brightness make detailed studies of them difficult. Because dwarfs are so common, any galaxy evolution/formation theory must clearly be able to predict and describe the properties of these galaxies. The luminosity functions of nearby galaxies in all environments reveal that dwarf galaxies are far more common than brighter galaxies. Within galaxy clusters there furthermore appears to be an overdensity of dwarf galaxies when compared to the field. The origin of these extra dwarfs, or if this excess is even real, remains unknown. The cluster luminosity function itself is not universal, but is strongly dependent on environment \\citep{sab03}, with the luminosity function often steeper in the more diffuse outer regions of clusters than in the denser cluster cores. The luminosity function also depends on the individual cluster, with each one having a characteristic luminosity function. For example, the Fornax Cluster is compact and rich in early-type galaxies, with a flat luminosity function faint-end slope of $\\alpha = -1.1$ \\citep{mie}. The Virgo Cluster has a high abundance of spiral galaxies and has a steeper luminosity function with a faint end slope of $\\alpha = -1.6$ (\\citealt{trenth02,sab03}). We know that both Local Group (LG) dwarf elliptical and dwarf spheroidal galaxies display varying star formation histories, with metal poor populations as old as the classical halo globular galaxies, but with evidence for star formation as little as 2-3 Gyr ago. Some low-mass LG galaxies such as Sagittarius, contain surprisingly high metal rich stellar populations considering their luminosities, as is also seen in clusters of galaxies. Previous spectroscopic and ground based imaging has revealed that dwarfs in the core of clusters are not a simple homogeneous population, with cluster core galaxies fainter than M$_{\\rm B}$ $= -15$ containing a mix of metallicities and ages (\\citealt*{Po,RSMP,c1}). The formation scenarios for dwarf galaxies fall into two categories. The first of these is that dwarfs are old, primordial objects (hierarchical model) (e.g. \\citealt{wf}). The second is that they have recently evolved or transformed from a progenitor galaxy population (e.g. \\citealt*{moore}). Within hierarchical models of galaxy formation, dwarf galaxies are formed from small density fluctuations in the early universe. The lowest mass systems (i.e. dwarf galaxies) form first, and massive galaxies are built from these in mergers. If star formation occurred soon after the gravitational collapse of initial density perturbations, dwarf galaxy halos would be amongst the first objects to form (e.g. \\citealt{ds}). Theory also predicts that such systems would form first in the densest environments, although cluster dEs would not necessarily contain the first stars formed in the universe. These halos could have formed their stars quite recently, if star formation was suppressed in some way \\citep{tully}. Dwarfs in clusters, on the other hand, could also be the remnants of stripped discs or dwarf irregulars. Through galaxy harassment and interactions, these progenitors become morphologically transformed into dEs. Cluster galaxies become stripped of their interstellar medium, and become dynamically heated by high-speed interactions with other galaxies and the gravitational potential of the cluster. To compensate, the galaxy loses stars, and over time the spiral can morphologically transform into a dE \\citep*{c3}. This model is supported by recent observations of embedded discs in dEs in the Virgo and Fornax clusters (\\citealt*{Bar,DeR}). Distant clusters at z $\\approx$ 0.4 are also filled with small spiral galaxies, but this population is largely absent in nearby clusters, where dwarf spheroidals make up the faint-end of the luminosity function \\citep{moore}. Downsizing \\citep{DeL} is the observed trend that star formation occurs later, and over more extended timescales, in smaller galaxies. In this scenario, dEs and lower mass galaxies formed or entered clusters after the giant galaxies. Low mass galaxies on average end their star formation after the giant elliptical galaxies. For example, the faint end of the red sequence in clusters is not formed until z $<$ 0.8 \\citep{DeL}, implying the luminosity weighted stellar populations of lower mass galaxies are younger than the stars in the giant elliptical galaxies. This seems to contradict the hierarchical method of galaxy formation, where dwarf systems form first. Also, dwarf elliptical galaxies are preferentially found in dense cluster environments, with few examples of isolated dEs, again contradicting the hierarchical model of galaxy formation. However, some dEs in clusters contain old stellar populations, so it is likely that multiple formation methods are needed to explain the origin of cluster dwarfs. The colour-magnitude relation (CMR) for cluster galaxies is well defined at bright magnitudes and forms a tight sequence. However, it is unclear what the shape of the red sequence is at fainter magnitudes \\citep{an}. \\citet*{c2} find that galaxies at the brightest 4 magnitudes of Perseus obey the CMR, with the fainter candidate members showing significant scatter from the relation. In contrast, \\citet{an} and others find no deviation from the CMR at fainter magnitudes in other clusters. Instead, the faint red sequence is an extrapolation of what is observed at brighter magnitudes. With spectroscopy, we can test this directly by finding confirmed dwarf members of the cluster to establish the true form of the colour magnitude relation. Spectroscopy can also be used to measure the faint-end of the luminosity function in galaxy clusters, which is fundamental in describing the galaxy population. The luminosity function contains important information on the formation and evolution of galaxies, and at low redshifts contains the combined influence of the galaxy initial mass function, and the effects of any evolutionary processes that have taken place in the cluster since its formation. In this paper, we use deep Keck spectroscopy to determine cluster membership for dwarfs at M$_{\\rm B}$ $<$ $-16$, and present the ages and metallicities of these galaxies based on the strengths of Balmer and metal absorption lines in their spectra. We confirm that twelve dwarfs are Perseus Cluster members, while the remaining objects are background galaxies. Using these results, we examine the colour-magnitude relation for the Perseus Cluster down to M$_{\\rm B}$ $= -12.5$. We also measure the luminosity function faint-end slope $\\alpha$ based on our cluster membership results. Our main conclusion is that some dwarf galaxies in Perseus have old, metal poor populations, whilst others are younger, metal rich systems. This suggests that dwarf galaxies in the core of the Perseus Cluster are not a simple, homogeneous population, but require multiple scenarios for their formation. This paper is organised as follows: in $\\S$ 2 we discuss the observations and the selection criteria for the dwarf galaxies, $\\S$ 3.1 identifies the cluster members, $\\S$ 3.2 presents the colour magnitude relation, and the central luminosity function is presented in $\\S$ 3.3. In $\\S$ 3.4 we derive luminosity weighted ages and metallicities for the newly confirmed cluster members. A discussion of the main results is presented in $\\S$ 4 and these results are summarised in $\\S$ 5. We assume the distance to the Perseus Cluster is 77 Mpc throughout this paper. ", "conclusions": "We have analysed Keck/LRIS spectra for twenty-three dwarf galaxy candidates in the Perseus Cluster, from which we have confirmed cluster membership for twelve systems based on radial velocities measured from absorption lines. We extend the confirmed member colour-magnitude relation for Perseus down to M$_{\\rm B}$ $= -12.5$, finding that the slope of the colour-magnitude relation becomes bluer when the low-surface brightness dwarf galaxies are included. The fainter dwarfs also scatter more from the colour-magnitude relation, following the trend observed by \\citet{RS} for low-mass galaxies in Fornax and Coma. This scatter can be interpreted as a spread in the metallicity distributions of dwarf galaxies, which has been inferred for dwarf galaxies in other clusters by \\citet{RSMP} and \\citet{Po}. After removing non-members from the $B$-band luminosity function for the Perseus Cluster we find a faint-end slope $\\alpha$ $= -1.26$ $\\pm$ 0.06, similar to the field. Previous studies of other galaxy clusters have found that the dwarf-to-giant ratio is a function of local projected galaxy density. Extending this study to the outer regions of Perseus would enable us to see how the luminosity function changes with the local galaxy density. Colours, morphologies, and central surface brightnesses are not sufficient criteria to confirm cluster membership, as this work has shown. Cluster membership cannot be confirmed without spectroscopy, so the faint-end luminosity functions calculated for galaxy clusters where membership has not been confirmed spectroscopically are most likely artificially steepened by non-members resulting in higher dwarf-to-giant ratios. By comparing the observed dwarf spectral absorption indices with population synthesis models of \\citet{tho}, we derive luminosity-weighted ages and metallicities for the dwarf galaxies. A range of ages is observed, ranging from older than 8 Gyr to younger than 5 Gyr. The metallicity distribution of the faint cluster members is also not that of a simple homogeneous population, with the younger galaxies typically having higher metallicities. More observations of dwarf galaxies in rich, nearby cluster environments are required in order to help improve our understanding of the formation scenarios for cluster dwarfs. Further spectroscopic work with a larger sample would also allow better constraints on the ages and metallicities of cluster dwarf galaxies and would help with modelling the formation of such galaxies." }, "0710/0710.2793_arXiv.txt": { "abstract": " ", "introduction": "SGRs are galactic X-ray stars that emit, during sporadic times of high activity, a large number of short-duration (around 0.1 s) bursts of hard X-rays (Duncan and Thompson, 1992). A SGR is thought to be a magnetar, being a strongly magnetized neutron star powered by a very strong magnetic field ($\\ge$ 10$^{15}$ Gauss). On 27 December 2004 a powerful burst of X- and $\\gamma$-rays from one of the most highly magnetized neutron stars (SGR 1806-20) of our Galaxy reached the Earth's environment (Hurley et al., 2005). The Solar system received a shock, which is thought to be due to a cataclysm in the magnetar that caused it to emit as much energy in two-tenths of a second as the Sun gives off in 250,000 years. The signature of this event on the Earth's magnetic field has not previously been investigated. Here, we present the first results of the magnetar footprints on magnetic data recorded by near-Earth satellites. The magnetar SGR 1806-20 is the third such event ever recorded along with two others that were noted in 1979 and 1998 (Mazets et al., 1979; Hurley et al., 1999). Several properties of this magnetar flare are relevant to our study. Firstly, a precursor of $\\sim$ 1 s was observed 142 s before the flare, with a roughly flat-topped profile (Hurley et al., 2005). The intensity of the main initial spike saturated all X- and $\\gamma$-ray detectors. However, particle detectors on board of RHESSI and Wind spacecraft (Boggs et al., 2004; Mazets et al., 2004) were able to record reliable measurements. Several instruments designed for other purposes provided important information, as Geotail (Terasawa et al., 2005) and Cluster/Double star (Schwartz et al., 2005). The first spike was followed by a tail lasting 380 s, during which 7.56 s pulsations were clearly observed, by the $\\gamma$-ray detectors on board of RHESSI (Hurley et al., 2005). Secondly, a disturbance of the Earth's ionosphere was simultaneously observed with the detection of the burst from SGR 1806-20 (Inan et al., 2005). This sudden ionospheric disturbance (SID) was recorded as a change in the signal strength from very low frequency (VLF) radio transmitters, being noticed by stations around the globe (Campbell et al., 2005). These changes in the radio signal strength were caused by X-rays arriving from SGR 1806-20, which ionized the upper atmosphere and modified the radio propagation properties of the Earth's ionosphere (see clearing house of SID data associated with SGR 1806-20 flare at http://www.aavso.org/observing/programs /solar/sid-sgr1806.shtml). One such observation of this ionospheric signature resides within a 21.4 kHz signal that originates in Hawaii and propagates along an ionosphere wave guide to Palmer Station, Antarctica (Inan et al., 2005). This wave guide is some $\\sim$ 10,000 km in path length (Inan et al., 2005). As explained above, this is not a direct radio detection of SGR 1806-20 (see also http://gcn.gsfc.nasa.gov/gcn3/2932.gcn3). Moreover, due to the sub-burst longitude and latitude (Inan et al., 2005) and to the geographical distribution of LF/VLF beacons and monitoring stations, this burst was not detected by active monitoring stations in Germany, Australia, or Canada (Campbell et al., 2005). Here, we note that ionospheric disturbances were also reported in the case of the magnetar observed in 1998 (Inan et al., 1999). In the case of the 1998 magnetar the flare illuminated the nightside of the Earth and ionized the lower ionosphere to levels usually found only during daytime. The magnetar responsible for the 2004 burst was about the same distance as the magnetar responsible for the 1998 burst, but within 5.25$^\\circ$ of the Sun as viewed from Earth. Therefore its $\\gamma$-rays arrived on the dayside of our planet. The 2004 flare changed the ionic density at an altitude of 60 km by six orders of magnitude (Inan, 2006). It is thus plausible that this change in the ionospheric conductivity can cause oscillating perturbations in the current-generated magnetic field. The thrust of this study is to find signatures associated with the explosion of the magnetar SGR 1806-20 within satellite measurements of Earth's electromagnetic field. Currently, the Earth's electromagnetic field is monitored by a number of Low Earth Orbit (LEO) satellite missions. After the launch of \\O rsted satellite in 1999, the knowledge of the near-Earth electromagnetic field has been dramatically improved (Hulot et al., 2002; L\\\"uhr et al., 2002; Maus et al., 2002; Tyler et al., 2003; Balasis et al, 2004). Since 2000, \\O rsted, CHAMP and SAC-C satellites have offered a continuous flow of high quality magnetic field measurements. Additionally, the DEMETER satellite provides 1 Hz energetic electron detector data. Finally, let us note that all these LEO magnetic missions are flying between the Earth's surface, where the temporal variations of the magnetic field are continuously monitored by geomagnetic observatories, and the magnetosphere, where an in-situ investigation of the three-dimensional and time-varying phenomena is done by the four identical spacecraft of Cluster II mission. ", "conclusions": "The effect of the SGR1806-20 flare on the Earth's magnetic field was not large, but it was detectable. This first attempt to find a magnetar signature in the geomagnetic field clearly indicates that the high resolution CHAMP magnetic data are optimal to capture the extremely bright flare from SGR 1806-20. Indeed, during the first half of the decay phase of the flare a 7.5 s periodicity is observed in the magnetic field over a magnetically quiet period, near the South Pole at 400 km altitude. This observation can be explained by a mechanism through which the oscillating flux of ionizing $\\gamma$-rays could alter the ionospheric conductivity and hence cause oscillating perturbations in the current-generated magnetic field. An attempt to verify this hypothesis for the two previously recorded giant flares was not possible since no magnetic satellite missions were operating in LEO at that time (i.e., in March 1979 and August 1998). Of course, there are many spacecraft carrying magnetometers within the Solar system, but very few near planetary ionosphere. For example, the wavelet analysis was performed on magnetometer data from the Cluster II mission that probes the Earth's magnetosphere. In order to be able to visualize, in the wavelet power spectrum graph, any significant disturbances of the magnetic field, the power spectral density of the signal was amplified by a factor of $2^6$ (in comparison to the corresponding spectral density values of the CHAMP data). Although there are some indications for a weak pulsation-like signal at $\\sim$ 8 s, the fact that this signal is almost two orders of magnitudes weaker than the one observed in CHAMP data favors the hypothesis of an ionospheric origin for the signature found in CHAMP data. Furthermore, our analysis can be extended to 1 Hz magnetic data provided by ground-based magnetic observatories, but only a small number of them provide such high resolution sampling nowadays. Data provided by 12 Canadian observatories, for which 1 Hz values are available over the period we are interested in, were also analyzed. For five of these observatories, missing data or high-level noise, made it difficult to apply the wavelet technique. For the others, no conclusive evidence for a signature related to SGR 1806-20 exists. Analyzing other magnetic data with such a powerful tool as wavelets techniques could be relevant for understanding the impact that giant flares have on the terrestrial and other planetary magnetic fields. However, the main difficulty in such studies is due to the availability and quality of magnetic data. For instance, wavelet analysis of Mars Global Surveyor mission magnetic measurements on 27/12/04 was not able to detect any of the magnetar features due to the inadequate sampling rate: only 3 s data are now available (Michael Purucker, pers. comm. 2005)." }, "0710/0710.3795_arXiv.txt": { "abstract": "Massive black hole binary coalescences are prime targets for space-based gravitational wave (GW) observatories such as {\\it LISA}. GW measurements can localize the position of a coalescing binary on the sky to an ellipse with a major axis of a few tens of arcminutes to a few degrees, depending on source redshift, and a minor axis which is $2 - 4$ times smaller. Neglecting weak gravitational lensing, the GWs would also determine the source's luminosity distance to better than percent accuracy for close sources, degrading to several percent for more distant sources. Weak lensing cannot, in fact, be neglected and is expected to limit the accuracy with which distances can be fixed to errors no less than a few percent. Assuming a well-measured cosmology, the source's redshift could be inferred with similar accuracy. GWs alone can thus pinpoint a binary to a three-dimensional ``pixel'' which can help guide searches for the hosts of these events. We examine the time evolution of this pixel, studying it at merger and at several intervals before merger. One day before merger, the major axis of the error ellipse is typically larger than its final value by a factor of $\\sim 1.5-6$. The minor axis is larger by a factor of $\\sim 2-9$, and, neglecting lensing, the error in the luminosity distance is larger by a factor of $\\sim 1.5-7$. This large change over a short period of time is due to spin-induced precession, which is strongest in the final days before merger. The evolution is slower as we go back further in time. For $z = 1$, we find that GWs will localize a coalescing binary to within $\\sim 10\\ \\mathrm{deg}^2$ as early as a month prior to merger and determine distance (and hence redshift) to several percent. ", "introduction": "\\label{sec:intro} Among the most important sources of gravitational waves (GWs) in the low-frequency band of space-based detectors are the coalescences of massive black hole binaries (MBHBs). Binaries containing black holes with masses in the range $10^4 - 10^7\\,M_\\odot$ are predicted to form through the hierarchical growth of structure as dark matter halos (and the galaxies they host) repeatedly merge; see, for example, {\\citet{svh07}} and {\\citet{mhsa07}} for recent discussion. The {\\it Laser Interferometer Space Antenna} ({\\it LISA}) is being designed to have a sensitivity that would allow detailed measurement of the waves from these binaries. ``Intrinsic'' parameters --- the masses and spins of the black holes which compose the binary --- should be determined with very high accuracy, with relative errors typically $\\sim 10^{-3}$ to $10^{-1}$, depending on system mass and redshift; see Lang \\& Hughes (2006, hereafter Paper I) for recent discussion. By measuring an ensemble of coalescences over a range of redshifts, MBHB GWs may serve as a kind of structure tracer, tracking the growth and spin evolution of black holes over cosmic time. ``Extrinsic'' system parameters, describing a binary's location and orientation relative to the detector, are also determined by measuring its GWs. In Paper I, we showed that a binary's position on the sky can be localized at $z = 1$ to an ellipse with a major axis of a few tens of arcminutes and a minor axis a factor of $2-4$ smaller. At higher redshift ($z = 3-5$), these values degrade by a factor of a few, reaching a few degrees in the long direction and tens of arcminutes to a degree or two in the short one. We also found that, neglecting errors due to weak gravitational lensing, a source's luminosity distance typically can be determined to better $1\\%$ at low redshift ($z = 1$), degrading to several percent at higher redshift ($z = 3-5$). The intrinsic ability of GWs to determine the distance to a coalescing binary is phenomenal. Coalescing MBHB systems constitute exquisitely well calibrated distance measures, with the calibration provided by general relativity. Unfortunately, this percent-level or better accuracy could only be achieved if we measured MBHB coalescences in an empty universe. In our universe, weak lensing will magnify or demagnify the GWs, and we will infer a luminosity distance smaller (for magnification) or larger (for demagnification) than the true value. This phenomenon affects all high-redshift standard candles. Its impact on Type Ia supernovae in particular has been discussed in detail {\\citep{frieman97,hw98,h98}}. It will not be possible to correct for this effect {\\citep{dhcf03}} since much of the lensing ``noise'' arises from structure on subarcminute scales that is not probed by shear maps (which map the distribution of matter on scales greater than $1^\\prime$ or so). Since we will not know the extent of the magnification when we measure MBHB waves, we must simply accept the fact that lensing introduces a dispersion of several percent in determining the distance to these GW events (see, e.g., Wang et al. [2002] to compute this dispersion as a function of redshift). When we quote distance measurement errors, we will typically quote only the intrinsic GW measurement error, neglecting lensing's impact. When the intrinsic GW distance error is $\\lesssim 5\\%$, lensing will blur it to the several percent level. Note that a source's redshift $z$ {\\it cannot} be directly determined using only GWs. Gravitational wave measurements infer system parameters through their impact on certain dynamical timescales, such as orbital frequencies and the rate at which these frequencies evolve. Since these time scales all suffer cosmological redshift, $z$ is degenerate with other parameters. For example, any mass parameter $m$ is actually measured as $(1 + z)m$. However, if the binary's luminosity distance is determined, its redshift can then be inferred by assuming a cosmography. For most binaries, the redshift can be determined to several percent (with an error budget dominated by gravitational lensing\\footnote{At redshifts $z \\lesssim 0.3$, the error is actually dominated by peculiar velocity effects {\\citep{kfhm06}}; however, the event rate is probably negligible at such low redshifts. As such, we will focus on gravitational lensing as the main source of systematic redshift error.}). We thus expect that GW measurements will locate a binary to within a three-dimensional ``GW pixel'' which at $z = 1$ has a cross-sectional area of $\\sim 10^{-2}$ to $10^{-1}\\ \\mathrm{deg}^2$ and a depth $\\Delta z/z \\sim $ several percent. It is anticipated that there will be great interest in searching the GW pixel for electromagnetic (e.g., optical, X-ray, radio) counterparts to MBHB GW events. Finding such a counterpart would be much easier if galactic activity were catalyzed in association with the coalescence {\\citep{kfhm06}}. The nature of that activity is likely to depend rather strongly on the mass of the coalescing system \\citep{dssch06}. For example, {\\citet{an02}} predict strong outflows and galactic activity prior to the final black hole merger as the smaller member of the binary drives gas onto the larger member, consistent with the high-mass ($M_{\\rm tot} \\gtrsim 10^7\\,M_\\odot$) predictions of \\citet{dssch06}. {\\citet{mp05}} describe an X-ray afterglow that would ignite after gas refills the volume that is swept clean by the coalescing binary; Dotti et al.\\ predict this outcome for smaller systems ($M_{\\rm tot} \\lesssim \\mbox{several}\\times 10^6\\,M_\\odot$). Recent work by {\\citet{bp07}} suggests that the final burst of radiation from a coalescing binary (which can convert $\\sim 10\\%$ of the system's mass to GWs very suddenly) may excite radial waves, and consequently electromagnetic variability, in an accretion disk due to the quick change in the disk's Keplerian potential. Such a signature may be essentially mass independent. On the other hand, the coalescence may be electromagnetically quiet, in which case we face the potentially daunting task of searching the three-dimensional pixel for galaxies with morphology consistent with a (relatively) recent merger, or that have a central velocity dispersion $\\sigma$ consistent with the inferred final black hole mass (assuming that the $M_{\\rm BH}-\\sigma$ relation [Ferrarese \\& Merritt 2000; Gebhardt et al. 2000] holds at the redshift of these sources, and so soon after merger). If the host galaxy or some other counterpart can be identified, we could then contemplate combining GW information with electromagnetic data. For instance, combining {\\it LISA} mass measurements with the luminosity of the counterpart may allow us to directly measure the Eddington ratio $L/L_{\\mathrm{Edd}}$ {\\citep{kfhm06}}. Identifying a counterpart would also allow us to more accurately characterize the system. For example, much of the intrinsic luminosity distance error is due to correlations between distance and sky position. Finding an electromagnetic counterpart essentially determines a binary's position precisely, breaking those correlations. Previous studies have found that intrinsic distance error can be reduced by almost an order of magnitude if the position is known {\\citep{h02,hh05}}. (Lensing errors still dominate in such a case, so that the distance remains determined only at the few percent level.) A counterpart may also make it possible to simultaneously determine a source's luminosity distance and redshift. Such a ``standard siren'' (the GW analog of a standard candle) would very usefully complement other high-redshift standard candles {\\citep{hh05}}, such as Type Ia supernovae {\\citep{p93, rpk95, wgap03}}. A direct measurement of redshift will also break the mass-redshift degeneracy more accurately than can be done with just the luminosity distance and some assumed cosmological parameters. Breaking this degeneracy is critical when studying the growth of black holes with cosmic time {\\citep{h02}}. Many analyses {\\citep{c98,h02,v04,bbw05,hh05}} have quantified how well {\\it LISA} can determine MBHB parameters, including sky position and distance, using maximum likelihood Fisher matrix estimation. Our results from Paper I, given earlier, include the effects of ``spin-induced precession'' --- precession of both the orbital plane and the individual spins of the black holes due to post-Newtonian spin-interaction effects. A significant result from that analysis is that spin-induced precession improves sky position accuracy by about half an order of magnitude in each direction versus previous analyses. This result can be partially understood as due to the breaking of a degeneracy between position and orientation angles: thanks to precession, the binary's orientation evolves with time and can be untangled from sky position. This effect was already known due to pioneering work by {\\citet{v04}}; by taking the analysis to higher order and considering a broader range of sources, we were able to show that this improvement held for essentially all astrophysically interesting MBHB sources. The purpose of this paper is to examine the localization of MBHB systems more thoroughly, in particular how the GW pixel evolves as the final merger is approached. Paper I only presented results for measurements that proceed all the way to merger. It will clearly be of some interest to monitor potential hosts for the binary event some time before the merger happens; if nothing else, telescopes will need prior warning to schedule observing campaigns. Understanding the rate at which localization evolves can also have an important impact on the design of the {\\it LISA} mission, clarifying how often it will be necessary to downlink data about MBHB systems in order to effectively guide surveys. Our main goal is to understand for what range of masses and redshifts prior localization of a binary using GWs will be possible. A previous analysis by Kocsis et al.\\ (2007b; hereafter K07) also examined this problem in great detail, but without including the impact of spin-induced precession. One of our goals is to see to what extent precession physics changes the conclusions of K07. We find that precession has a fairly small impact on the time evolution of the GW pixel except in the last few days before the final merger, at which point its impact can be tremendous. Precession typically changes the area of the sky position error ellipse by a factor of $\\sim 3-10$ (up to $\\sim 60$ in extreme cases) in just the final day. This is in accord with the predictions of K07 (and even earlier predictions by N. Cornish 2005, unpublished). The structure of this paper is as follows. First, in \\S\\ {\\ref{sec:background}}, we briefly review the basics of the MBHB gravitational waveform and the parameter estimation formalism that we use; this section is essentially a synopsis of relevant material from Paper I. Section {\\ref{sec:intrinsicGW}} reviews the form of the GWs that we use in our analysis, while \\S\\ {\\ref{sec:extrinsicGW}} describes how those waves are measured by the {\\it LISA} constellation. We describe the measurement formalism we use in \\S\\ {\\ref{sec:formalism}}. In \\S\\ {\\ref{sec:review}}, we summarize our results from Paper I regarding the final localization accuracy that {\\it LISA} can expect to achieve. We turn to a detailed discussion of the time evolution of the GW pixel in \\S\\ {\\ref{sec:results}}. We begin by summarizing the key ideas behind the ``harmonic mode decomposition'' of K07 in \\S\\ {\\ref{sec:khmf}}. This technique cleverly allows calculation of the GW pixel and its time evolution with much less computational effort than our method (albeit without including the impact of spin precession). Unfortunately, we have discovered that some of the approximations used by K07 introduce a systematic underestimate of the final sky position error by a factor of $2 - 4$ or more in angle; the approximations are much more reliable a week or more prior to the black holes' final merger. Modulo this underestimate, the K07 results agree well with a version of our code which does not include spin precession (particularly a week or more in advance of merger, when their underestimate is not severe). K07 thus serves as a useful point of comparison to establish the impact of precession on source localization. Section {\\ref{sec:timeevolve}} is dedicated to our results, including comparison to K07 when appropriate. We find that all relevant parameter errors decrease slowly with time until the last day before merger, when they drop more dramatically. This sudden drop is not found in K07, nor is it present in a variant of our analysis that ignores spin precession. It clearly can be attributed to the impact of precession on the waveform. Before this last day, the major axis is $\\sim 1.5-6$ times, the minor axis $\\sim 2-9$ times, and the intrinsic error in the luminosity distance $D_L$ $\\sim 1.5-7$ times bigger than at merger for most binaries (i.e., all except the highest masses). Going back to one week (one month) before merger, these numbers change to $2-9$ ($4-11$) for the major axis, $3-14$ ($5-24$) for the minor axis, and $3-14$ ($5-18$) for the error in the luminosity distance. As a result, for $z = 1$, most binaries can be located within a few square degrees a week before merger and $10\\ \\mathrm{deg}^2$ a month before merger. The intrinsic distance errors are also small enough this early that $\\Delta z/z$ remains dominated by gravitational lensing errors of several percent. Advanced localization of MBHB coalescences thus seems plausible for these binaries; the situation is less promising for sources at higher redshift. As a corollary to our study of the time evolution, we also examine the sky position dependence of errors (\\S\\ {\\ref{sec:angdependence}}). The errors depend strongly on the polar angle with respect to the ecliptic, increasing in the ecliptic plane to as much as $35\\%$ over the median for the major axis, $85\\%$ over the median for the minor axis, and $15\\%$ over the median for errors in the luminosity distance. The errors have a much weaker dependence on the azimuthal angle. When we convert to Galactic coordinates, we find that the best localization regions appear to lie fairly far out of the Galactic plane, offering hope that searches for counterparts will not be too badly impacted by foreground contamination. We conclude this paper in \\S\\ {\\ref{sec:disc}}. Besides summarizing our results, we discuss shortcomings of this analysis and future work which could help to better understand how well GWs can localize MBHB sources. Throughout the paper we set $G = c = 1$; a convenient conversion factor in this system is $10^6 M_\\odot = 4.92$ s. When discussing results, we always quote masses as they would be measured in the rest frame of the source. These masses must be multiplied by a factor of $1 + z$ when used in any of the equations describing the system's dynamics or its GWs (particularly the equations of \\S\\ {\\ref{sec:background}}). We convert between distance and redshift using a spatially flat cosmology with $\\Omega_\\Lambda = 0.75$ (and hence $\\Omega_m = 0.25$) and Hubble constant $H_0 = 75\\ {\\rm km}\\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1}$. ", "conclusions": "\\label{sec:disc} As discussed at length in Paper I, accounting for the general relativistic precession of the angular momentum vectors in an MBHB system has a dramatic impact on what we can learn by observing the system's gravitational waves. Spin-induced precession breaks degeneracies among different parameters, making it possible to measure them more accurately than they could be determined if precession were not present. This has a particularly important impact on our ability to locate such a binary on the sky and to determine its luminosity distance --- the degeneracy between sky angles, distance, and orientation angles is severe in the absence of precession. Our analysis shows that the improvement that precession imparts to measurement accumulates fairly slowly. In using one code which includes the impact of spin precession and a second which neglects this effect, we find little difference in the accuracy with which GWs determine sky position and distance for times more than a few days in advance of the final merger. The difference between the two codes grows quite rapidly in these final days. In the last day alone, the localization ellipse area decreases by a factor of $\\sim 3-10$ (up to $\\sim 60$ in a few low-mass systems) when precession effects are included. Distance determination is likewise improved by factors of $\\sim 1.5-7$ in that final day. Not all of the precession effects occur in the final days. We saw in Figure \\ref{fig:axes_evol} that the long tail of small minor axes can be seen, to some degree, throughout the inspiral. We could get lucky and find a binary with a very small value of $2b$ weeks before merger. But the improvement in the median that we found in Paper I appears to take effect only in the final days of inspiral. Therefore, while precession may in fact help improve the {\\it final} localization of a coalescing binary by a factor of $\\sim 2-10$ in each direction, it will not be much help in {\\it advanced} localization of a typical binary. Nevertheless, the pixel sizes that we find are small enough that future surveys should not have too much trouble searching the region identified by GWs, at least over certain ranges of mass and redshift. At $z = 1$, the GW localization ellipse is $\\sim 10\\ \\mathrm{deg}^2$ or smaller for most binaries as early as a month in advance of merger. (At high masses, the ellipse can be substantially larger than this a month before merger, but it shrinks rapidly, reaching a comparable size $1-2$ weeks before merger.) This bodes well for future surveys with large fields of view that are likely to search the GW pixel for counterparts. In addition, GWs determine the source luminosity distance so well that the distance errors we find are essentially irrelevant --- gravitational lensing will dominate the distance error budget for all but the highest masses. As redshift increases, the GW pixel rapidly degrades, particularly for the largest masses. Let us adopt $10\\ \\mathrm{deg}^2$ (the approximate LSST field of view) as a benchmark localization for which counterpart searches may be contemplated. At $z = 3$, this benchmark is reached at merger for almost the entire range of masses we considered. As little as a day in advance of merger, however, some of the least massive and most massive systems are out of this regime. One week prior to merger, the most massive systems are barely located at all (ellipses hundreds of square degrees or larger). The intermediate masses do best, but even in their cases the positions are determined with $\\sim 10\\ \\mathrm{deg}^2$ accuracy no earlier than a few days in advance of merger. The resolution degrades further at higher redshift. At $z = 5$, systems with $M \\gtrsim 6 \\times 10^6\\,M_\\odot$ are not located more accurately than $\\sim 30\\ \\mathrm{deg}^2$ even at merger. Smaller systems are located within $\\sim 10\\ \\mathrm{deg}^2$ at merger, but very few are at this accuracy even one day in advance of merger. The luminosity distance errors also increase, so much that they exceed lensing errors a few days to a week before merger at $z = 3$, and only a day before merger at $z = 5$. This degradation hurts the ability to search for counterparts by redshift and subsequently use them as standard candles. Our main conclusion is that future surveys are likely to have good advanced knowledge (a few days to one month) of the location of MBHB coalescences at low redshift ($z \\sim 1 - 3$), but only a day's notice at most at higher redshift ($z \\sim 5$). This conclusion may be excessively pessimistic. As mentioned earlier, recent work examining the importance of subleading harmonics of MBHB GWs is finding that including harmonics beyond the leading quadrupole has an important effect on the final accuracy of position determination {\\citep{arunetal,ts08}}. For most masses, these analyses show a factor of a few improvement in position, comparable to the improvement that we find when spin precession is added to the waveform model. For high-mass systems, the higher harmonics increase the (previously small) overlap with the {\\it LISA} band; consequently, the improvement can be much larger, up to 2 or 3 orders of magnitude in area. Since these two improvements arise from very different physical effects, it is likely that their separate improvements can be combined for an overall improvement significantly better than each effect on its own. We plan to test this in future work (which is just now getting underway). Finally, we have also studied the sky position dependence of {\\it LISA}'s ability to localize sources. We have found that the regions of best localization lie fairly far out of the Galactic plane. However, as emphasized by N. Cornish (2007, private communication), a proper anisotropic confusion background might impact this dependence. In our calculations, we have assumed an isotropic background, neglecting the likely spatial distribution of Galactic binaries. Properly accounting for this background is likely to strengthen our conclusion that LISA's ability to ``see'' is best for MBHB sources out of the Galactic plane." }, "0710/0710.3106_arXiv.txt": { "abstract": "Development of the Aarhus adiabatic pulsation code started around 1978. Although the main features have been stable for more than a decade, development of the code is continuing, concerning numerical properties and output. The code has been provided as a generally available package and has seen substantial use at a number of installations. Further development of the package, including bringing the documentation closer to being up to date, is planned as part of the HELAS Coordination Action. ", "introduction": "The goal of the development of the code was to have a simple and efficient tool for the computation of adiabatic oscillation frequencies and eigenfunctions for general stellar models, emphasizing also the accuracy of the results. Not surprisingly, given the long development period, the simplicity is now less evident. However, the code offers considerable flexibility in the choice of integration method as well as ability to determine all frequencies of a given model, in a given range of degree and frequency. The choice of variables describing the equilibrium model and oscillations was to a large extent inspired by \\citet{Dziemb1971}. As discussed in Section~\\ref{sec:eqmodel} the equilibrium model is defined in terms of a minimal set of dimensionless variables, as well as by mass and radius of the model. Fairly extensive documentation of the code, on which the present paper in part is based, is provided with the distribution package% \\footnote{The package is available at \\\\ {\\tt http://astro.phys.au.dk/$\\sim$jcd/adipack.n}}. \\citet{Christ1991} provided an extensive review of adiabatic stellar oscillations, emphasizing applications to helioseismology, and discussed many aspects and tests of the Aarhus package, whereas \\citet{Christ1994} carried out careful tests and comparisons of results on polytropic models; this includes extensive tables of frequencies which can be used for comparison with other codes. ", "conclusions": "" }, "0710/0710.2954_arXiv.txt": { "abstract": "We have compiled the emission-line fluxes of \\ion{O}{1} $\\lambda$8446, \\ion{O}{1} $\\lambda$11287, and the near-infrared (IR) \\ion{Ca}{2} triplet ($\\lambda$8579) observed in 11 quasars. These lines are considered to emerge from the same gas as do the \\ion{Fe}{2} lines in the low-ionized portion of the broad emission line region (BELR). The compiled quasars are distributed over wide ranges of redshift (0.06 $\\le z \\le$ 1.08) and of luminosity ($-29.8 \\le M_{B} \\le -22.1$), thus providing a useful sample to investigate the line-emitting gas properties in various quasar environments. The measured line strengths and velocities, as functions of the quasar properties, are analyzed using photoionization model calculations. % We found that the flux ratio between the \\ion{Ca}{2} triplet and \\ion{O}{1} $\\lambda$8446 is hardly dependent on the redshift or luminosity, indicating similar gas densities in the emission region from quasar to quasar. On the other hand, a scatter of the \\ion{O}{1} $\\lambda$11287/$\\lambda$8446 ratios appears to imply the diversity of the ionization parameter. These facts invoke a picture of the line-emitting gases in quasars that have similar densities and are located at regions exposed to various ionizing radiation fluxes. The observed \\ion{O}{1} line widths are found to be remarkably similar over more than 3 orders of magnitude in luminosity, which indicates a kinematically determined location of the emission region and is in clear contrast to the case of \\ion{H}{1} lines. We also argue about the dust presence in the emission region since the region is suggested to be located near the dust sublimation point at the outer edge of the BELR. ", "introduction": "Active galactic nuclei (AGNs) are known to have strong emission lines of various ion species. Among them, the \\ion{Fe}{2} emission lines are one of the most prominent features in the ultraviolet (UV) to optical spectrum of many AGNs. They have long been hoped to provide significant information about some aspects of the AGNs and their host environments, e.g., energy budget of the line emission region % and the epoch of the first star formation in the host galaxies. The determination of the first star formation epoch is based on the standard theory that the iron enrichment in galaxies is delayed compared to that of the $\\alpha$-elements, such as magnesium, due to their different origins; Type Ia supernovae for iron and Type II supernovae for the $\\alpha$-elements \\citep{hamann93,yoshii98}. The delay corresponds to the difference in life-times of the progenitors of the two types of supernovae, and is estimated to be 0.3 -- 1 Gyr depending on the host galaxy environments \\citep{yoshii96, matteucci01}. Many observations have been devoted to the measurement of \\ion{Fe}{2}/\\ion{Mg}{2} line flux ratios in high-redshift quasars for this purpose over the last decade \\citep[e.g.,][]{elston94,kawara96,dietrich02,dietrich03,iwamuro02,iwamuro04,freudling03,maiolino03}. However, the observed \\ion{Fe}{2}/\\ion{Mg}{2} ratios show a large scatter, preventing a detection of any significant trend in the Fe abundance as a function of redshift. While a part of the scatter might be due to the difference in the intrinsic Fe/Mg abundance ratio, it is presumable that the diversity of the physical condition within the line-forming gas, affecting line emissivities, is the main cause \\citep{verner03, baldwin04}. In the same sense, a change of the observed \\ion{Fe}{2}/\\ion{Mg}{2} ratio as a function of redshift, if found, should be carefully examined to tell whether it reflects the abundance evolution or the systematic variation of the line emissivity. Thus establishment of a method to probe the line-emitting gas and estimate its physical parameters such as density and incident-ionizing radiation flux has been much awaited. Unfortunately, the Fe$^+$ atom is characterized by an enormous numbers of possible electronic transitions, yielding the ``\\ion{Fe}{2} pseudocontinuum'' often observed in AGN spectra, which makes analysis of the observations extremely difficult from both the observational and theoretical viewpoints \\citep[e.g., ][hereafter T06]{tsuzuki06}. On the other hand, a promising approach is to use the emission lines emitted by simple atoms in the same region as the \\ion{Fe}{2} lines. The most potent lines are \\ion{O}{1} and \\ion{Ca}{2}, whose co-spatial emergence with \\ion{Fe}{2} is indicated by a resemblance of their profiles \\citep{rodriguez02a} and by a correlation between the line strengths \\citep{persson88}. Note that it is a natural consequence of similar ionization potentials of the relevant ions, i.e., 16.2 eV for \\ion{Fe}{2}, 13.6 eV for \\ion{O}{1}, and 11.9 eV for \\ion{Ca}{2}. The first extensive study of the physical properties of \\ion{O}{1} emitting gases in AGNs was presented by \\citet{grandi80}, who observed the strongest \\ion{O}{1} line, $\\lambda$8446, as well as other weaker \\ion{O}{1} lines in Seyfert 1 (Sy1) galaxies. He found that \\ion{O}{1} $\\lambda$8446 lacks the narrow component that characterizes other permitted lines, and concluded that the line is purely a BELR phenomenon. He also suggested that \\ion{O}{1} $\\lambda$8446 is produced by Ly$\\beta$ fluorescence, which was later confirmed by the observation of I Zw 1, the prototype narrow-line Seyfert 1 (NLS1), by \\citet{rudy89}. \\citet{rodriguez02b} compiled the UV and near-IR \\ion{O}{1} lines, namely, $\\lambda$1304, $\\lambda$8446, and $\\lambda$11287, in normal Sy1s and NLS1s in order to investigate their flux ratios. They found that there must be an additional excitation mechanism for \\ion{O}{1} $\\lambda$8446---besides Ly$\\beta$ fluorescence---which they concluded is collisional excitation. As for the \\ion{Ca}{2} lines, extensive studies of Sy1s were presented by \\citet{persson88} and \\citet{ferland89}. We have observed seven quasars at redshifts up to $\\sim$1.0 for the purpose of obtaining the UV and near-IR \\ion{O}{1} and \\ion{Ca}{2} lines, thus extending the previous studies to include quasars at high redshifts. Photoionization model calculations were performed and compared with the observations, which led us to conclude that the \\ion{O}{1} and \\ion{Ca}{2} lines are formed in a gas with density $n_{\\rm H} = 10^{11.5} - 10^{12.0}$ cm$^{-3}$, illuminated by the ionizing radiation corresponding to the ionization parameter of $U = 10^{-3.0} - 10^{-2.5}$ \\citep[][the latter is referred to as Paper I hereafter]{matsuoka05, matsuoka07}. Now that the general properties of the \\ion{O}{1} and \\ion{Ca}{2} emitting gas, as a whole, are thus being revealed, we should turn our eyes to the following question: how are these gas properties related to the quasar characteristics, such as redshift and luminosity? Here we present results of a compilation of the emission-line fluxes of \\ion{O}{1} $\\lambda$8446, \\ion{O}{1} $\\lambda$11287, and the near-IR \\ion{Ca}{2} triplet ($\\lambda$8498, $\\lambda$8542, and $\\lambda$8662) observed in 11 quasars. The quasars are distributed over wide ranges of redshift (0.06 $\\le z \\le$ 1.08) and of luminosity ($-29.8 \\le M_{B} \\le -22.1$), thus providing a useful sample to track the line-emitting gas properties in various quasar environments. The observational and theoretical data used in this work are described in \\S \\ref{sec:data}, and results and discussion appear in \\S \\ref{sec:results}. Our conclusions are summarized in \\S \\ref{sec:summary}. ", "conclusions": "} \\subsection{A Picture of the Dust-Free Emission Region} We plot the observed values of $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446), \\ion{O}{1} $n$($\\lambda$11287)/$n$($\\lambda$8446), and EW (\\ion{O}{1} $\\lambda$8446) as functions of the redshift and of the $B$-band absolute magnitude $M_{B}$ in Figure \\ref{vsprop2}. One of the remarkable results is found in the top panels; the \\ion{Ca}{2}/\\ion{O}{1} $\\lambda$8446 ratio is hardly dependent on redshift or luminosity over the plotted range, while the ratio is predicted to be very sensitive to the density of the line-emitting gas in the photoionization models. We show the model predictions on the $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446) -- \\ion{O}{1} $n$($\\lambda$11287)/$n$($\\lambda$8446) plane in Figure \\ref{oioi} ({\\it left}), as well as the observed values (see also Fig. 7 in Paper I\\footnote{ Note that Figure 7 in Paper I shows the predictions of Model 1, which is not adopted in this paper since the assumed microturbulent velocity, $v_{\\rm turb}$ = 0 km s$^{-1}$, could not reproduce the observed \\ion{Fe}{2} UV emissions (see \\S \\ref{data_model}). However, Model 3 adopted in this work predict very similar results to those shown in Figure 7 regarding the \\ion{O}{1} and \\ion{Ca}{2} emissions, while the whole pattern of contour is slightly ($\\sim$ 0.5 dex) shifted to the high-density regime; the best-fit parameters in Model 1 are ($n_{\\rm H}$, $U$) = (10$^{11.5}$ cm$^{-3}$, 10$^{-3.0}$). }). The gas density $n_{\\rm H}$ and the ionization parameter $U$ are changed around the reference grid point, ($n_{\\rm H}$, $U$) = (10$^{12.0}$ cm$^{-3}$, 10$^{-2.5}$), over 2 orders of magnitude in both parameters. It is clearly seen that the predicted $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446) ratio increases monotonically with the increased gas density, and that all the observed values are marked with the density in the vicinity of the reference point, log $n_{\\rm H}$ = 12.0. Thus the similar density of the line-emitting gases are strongly indicated for the quasars distributed over these redshift and luminosity ranges. The values of EW (\\ion{O}{1} $\\lambda$8446), both observed and calculated with the density of log $n_{\\rm H}$ = 12.0, are shown in Figure \\ref{oioi} ({\\it right}) as a function of the \\ion{O}{1} $n$($\\lambda$11287)/ $n$($\\lambda$8446) ratio. It shows that the predictions of EWs are also consistent with the observed data when log $n_{\\rm H}$ = 12.0 and a covering fraction (cf) of the line-emitting gas as seen from the central energy source of 0.2 -- 0.5 are assumed. Note that the covering fraction could be much smaller if we assumed oxygen overabundance relative to the solar value. On the other hand, a scatter of the observed data in the \\ion{O}{1} $n$($\\lambda$11287)/ $n$($\\lambda$8446) axis seems to be related to the diversity of the ionization parameter (Fig. \\ref{oioi}, {\\it left}). Note that the diversity of other parameters, such as microturbulent velocity, gas column density, and chemical composition, could not explain this diagram since they significantly alter the $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446) ratio, rather than \\ion{O}{1} $n$($\\lambda$11287)/ $n$($\\lambda$8446), and thus they are unable to explain the observed similarity of the former ratios (Paper I). It is also quite unlikely that the diversity of these parameter values is balanced out by the fine-tuned density in such a way that the $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446) ratio is always kept to be $\\sim$1.0, unless these lines are the dominant heating or cooling sources of the emission region. As with the $n$(\\ion{Ca}{2})/$n$(\\ion{O}{1} $\\lambda$8446) ratio, \\ion{O}{1} $n$($\\lambda$11287)/ $n$($\\lambda$8446) is not clearly dependent on the redshift or luminosity (Fig. \\ref{vsprop2}, {\\it middle left and middle right}). The above arguments invoke a picture of the line-emitting gases in quasars that have similar densities and are located at regions exposed to various ionizing radiation fluxes. It would be a consequence of the difference in distance to the central continuum source and/or in the intrinsic luminosity of the quasars. Note that it is in clear contrast to the well-studied case of H$\\beta$, whose emission regions in AGNs are known to be characterized by similar ionization parameters. In fact, reverberation mapping results for H$\\beta$ show the emission region size ($r$) -- luminosity ($L$) relation of $r \\propto L^{0.5}$, which is consistent with the constant ionization parameter regime \\citep{peterson02, bentz06}. Such a situation has long been expected in order to account for the remarkably similar AGN spectra over a broad range of luminosity, and was incorporated into the locally optimally emitting cloud (LOC) model suggested by \\citet{baldwin95}, that is, that the BELR is composed of gas with widely distributed physical parameters and each emission line arises from its preferable environment. On the other hand, the case in \\ion{O}{1} and \\ion{Ca}{2} lines apparently indicates that the location of the emission region is not radiation-selected. In line with the above arguments, we found a clear difference of the velocity-luminosity relation between \\ion{O}{1} and H$\\beta$; the measured \\ion{O}{1} line widths are plotted versus $M_B$ in Figure \\ref{mag_width}, which shows that the \\ion{O}{1} line widths are remarkably similar, concentrated around 1500 -- 2000 km s$^{-1}$, over more than 3 orders of magnitude in the $B$-band luminosity. On the other hand, those for the hydrogen Balmer lines usually have a large scatter as shown by, e.g., \\citet{kaspi00}; their sample of 34 AGNs, spanning over 4 orders of magnitude in continuum luminosity, has the line widths of 1000 -- 10,000 km s$^{-1}$. Such a trend is also indicated by \\citet{persson88}, who reported that while the correlation between FWHM (\\ion{O}{1}) and FWHM (H$\\beta$) is good for the small FWHM regime, the \\ion{O}{1} lines grow systematically narrower than H$\\beta$ at large line width. \\citet{rodriguez02a} conducted a detailed study of the near-IR emission line profiles in NLS1s, and found that \\ion{O}{1}, \\ion{Ca}{2}, and \\ion{Fe}{2} lines are systematically narrower than the broad components of other low-ionization lines such as hydrogen Paschen lines and \\ion{He}{1} $\\lambda$10830. They also argued that these lines are produced in the outermost portion of the BELR, since their widths are just slightly broader than those of [\\ion{S}{3}] $\\lambda$9531 which they assumed is formed in the inner portion of the narrow emission line region (NELR). While the scattered widths of \\ion{H}{1} lines could be interpreted as a consequence of the radiation-selected locations of the emitting gases, regardless of the gas kinematics, the remarkable similarity of the \\ion{O}{1} line widths might imply the kinematically determined emission regions. In such a situation, the diversity of the ionization parameters, as discussed above, would be an inevitable result. As stated by \\citet{persson88}, there is clearly interesting information which could be deduced from studies of the \\ion{O}{1} and \\ion{H}{1} line profiles; especially, reverberation mapping of these \\ion{O}{1} and \\ion{Ca}{2} lines would be a powerful tool to reveal the underlying physics. \\subsection{A Picture of the Dusty Emission Region} The widely-accepted theory of the BELR describes its outer edge, where the \\ion{O}{1} and \\ion{Ca}{2} emission lines are likely to be formed, set by the dust sublimation \\citep[e.g.,][]{laor93, netzer93}. If we accept this picture, the dust grains are possibly mixed in the line-emitting gas and suppress the \\ion{Ca}{2} emission through the substantial Ca depletion. % Such a situation is in fact reported for the NELR by the absence or significant weakness of the observed [\\ion{Ca}{2}] $\\lambda$7291 line \\citep{kingdon95, villar97}. \\citet{ferguson97} presented the LOC model calculations of the narrow emission lines and argued that Ca is depleted relative to the solar value by factors of 3 -- 160. It is hard to see an evidence of the dust presence in the emission region from our results, since Figure \\ref{oioi} appears to show that our dust-free models successfully reproduce the observations. However, it is noteworthy that % it is quite difficult to account for % the observed data at \\ion{O}{1} $n$($\\lambda$ 11287)/$n$($\\lambda$ 8446) $<$ 0.4 by the models with log $n_{\\rm H}$ = 12.0. The problem is that such small values of \\ion{O}{1} $n$($\\lambda$ 11287)/$n$($\\lambda$ 8446) could only be reproduced with the higher densities than log $n_{\\rm H}$ = 12.0 so that \\ion{O}{1} $\\lambda$8446 emission is exclusively enhanced by the collisional excitation, while such a dense gas produces intense \\ion{Ca}{2} emission that is much stronger than observed. One can clearly see this trend in Figure \\ref{oioi} ({\\it left}). If we assumed significant Ca depletion in the line-emitting gas, these difficulties are naturally resolved since it significantly suppresses the otherwise intense \\ion{Ca}{2} emissions. For example, the data point representing (log $n_{\\rm H}$, log $U$) = (13.0, $-$2.5) in Figure \\ref{oioi} ({\\it left}) could provide a plausible model for the observations at \\ion{O}{1} $n$($\\lambda$ 11287)/$n$($\\lambda$ 8446) $<$ 0.4 if the \\ion{Ca}{2} emission is suppressed by a factor of a few times 10.\\footnote{ The predicted EW for the model with (log $n_{\\rm H}$, log $U$) = (13.0, $-$2.5) is EW (\\ion{O}{1} $\\lambda$8446) = 30 \\AA, which is consistent with the observations if rather large covering factors of 0.5 -- 1.0 and/or oxygen overabundance relative to the solar value are assumed. } The dust grains present in the line-emitting gas might also give the natural explanation to the observed lack of some UV emission lines relative to their optical or near-IR counterparts. \\citet{ferland89} mentioned the possibility of the dust survival in the BELR gas in order to explain the extreme weakness of the observed \\ion{Ca}{2} $\\lambda$3934 and $\\lambda$3639 lines relative to the near-IR triplet. At least a part of the long-standing \\ion{Fe}{2} UV/opt problem, in which the observed ratios of \\ion{Fe}{2} UV flux to the optical flux fall far below the photoionization model predictions \\citep[see, e.g.,][]{baldwin04}, could also be explained. However, it should be noted that the dust would affect the line formation processes in a very complicated manner, which should be precisely addressed when discussing the specific lines; for the \\ion{O}{1} and \\ion{Ca}{2} lines, one of the most apparent effects as well as the Ca depletion would be the destruction of Ly$\\beta$ photons which otherwise excite \\ion{O}{1} atoms, thus the suppression of the \\ion{O}{1} emissions. The resultant line flux ratios could be much different from those derived from the simple speculations. } We have compiled the emission-line fluxes of \\ion{O}{1} $\\lambda$8446, \\ion{O}{1} $\\lambda$11287, and the near-IR \\ion{Ca}{2} triplet ($\\lambda$8579) observed in 11 quasars. The quasars are distributed over wide ranges of redshift (0.06 $\\le z \\le$ 1.08) and of luminosity ($-29.8 \\le M_B \\le -22.1$), thus providing a useful sample to track the line-emitting gas properties in various quasar environments. The measured line strengths and velocities, as functions of the quasar properties, were analyzed with photoionization model calculations. Our findings and conclusions are as follows: 1. There is no sign of a significant change in the flux ratios of the \\ion{Ca}{2} triplet and \\ion{O}{1} $\\lambda$8446 over the redshift and luminosity ranges studied here. It strongly indicates similar gas densities in the line-emission region from quasar to quasar. 2. The observed scatter of the \\ion{O}{1} $\\lambda$11287/$\\lambda$8446 ratios appears to be related to the diversity of the ionization parameter, while the ratio is not clearly dependent on the redshift or luminosity. Combined with the similarity of the \\ion{Ca}{2}/\\ion{O}{1} $\\lambda$8446 ratios, it invokes the picture of the line-emitting gases in quasars that have similar densities and are located at regions exposed to various ionizing radiation fluxes. 3. The \\ion{O}{1} line widths are remarkably similar from quasar to quasar over more than 3 orders of magnitude in luminosity. It might imply a kinematically determined location of the line emission region and is in clear contrast to the case of \\ion{H}{1} lines, whose emission region is considered to be radiation-selected. 4. If we accept that the \\ion{O}{1} and \\ion{Ca}{2} emission lines are formed at the outer edge of the BELR and that the outer edge is set by the dust sublimation, the line-emitting gas is possibly mixed with the dust grains. In fact such a situation may better reproduce the observations than the dust-free case through the significant Ca depletion." }, "0710/0710.5603_arXiv.txt": { "abstract": "We review the current status of accelerator, direct and indirect Dark Matter (DM) searches, focusing on the complementarity of different techniques and on the prospects for discovery. After taking a census of present and upcoming DM-related experiments, we review the motivations to go beyond an \"accelerator-only\" approach, and highlight the benefits of multidisciplinarity in the quest for DM. ", "introduction": "The evidence for non-baryonic dark matter is compelling at all observed astrophysical scales~\\cite{Bergstrom:2000pn,Bertone:2004pz}. Although alternative explanations in terms of modified gravity (see Ref.~\\cite{Bekenstein:2004ne} for a relativistic theory of the MOND paradigm) cannot be ruled out, they can hardly be reconciled with the most recent astrophysical observations~\\cite{Clowe:2006eq} without requiring additional matter beyond the observed baryons (e.g. Ref.~\\cite{Feix:2007zm} and references therein). It is therefore natural to ask {\\it how can we identify the nature of DM particles?}. We review here the main strategies that have been devised to attack this problem, namely accelerator, direct and indirect searches, focusing on the interplay between them and on their complementarity. In fact, a tremendous theoretical and experimental effort is in progress to clarify the nature of DM, mostly devoted, but not limited, to searches for Weakly Interacting Massive Particles (WIMPs), that achieve the appropriate relic density by {\\it freezing-out} of thermal equilibrium when their self-annihilation rate becomes smaller than the expansion rate of the Universe. The characteristic mass of these particles is $\\cal{O}$$(100)$ GeV, and the most representative and commonly discussed candidates in this class of models are the supersymmetric neutralino, and the B$^{(1)}$ particle, first excitation of the hypercharge gauge boson, in theories with Universal Extra Dimensions. A tentative census of present and upcoming DM experiments (WIMPs only) is shown in fig. 1. Shown in the figure are: two particle accelerators, viz. the Tevatron at Fermilab, and the upcoming Large Hadron Collider (LHC) at CERN; the many direct detection experiments currently taking data or planned for the near future, along with the names of the underground laboratories hosting them; high-energy neutrino telescopes; gamma-ray observatories; gamma-ray and anti-matter satellites. Light blue points denote gamma-ray experiments that are not directly related to indirect DM searches, as DM signals would be typically produced at energies below their energy threshold. Nevertheless, they may turn out to be useful to discriminate the nature of future unidentified high-energy gamma-ray sources. Three satellites are shown in the inset of figure 1: PAMELA, an anti-matter satellite that has already been launched and is expected to release the first scientific data very soon. ; GLAST, a gamma-ray satellite that is scheduled for launch in early 2008; and AMS-02, anti-matter satellite that should be launched in the near future. We will discuss below the prospects for detecting DM with the various experiments shown in fig. 1, and we will focus our attention on the complementarity of the various detection strategies. The paper is organized as follows: we first discuss accelerator searches, and show that although the LHC has the potential to make discoveries of paramount importance for our understanding of DM, it may not be able to solve all problems. In Section 3 we discuss the information that can be extracted from direct detection experiments, in case of positive detection. Section 4 is then dedicated to indirect searches, and to the question of what astrophysical observations can tell us about the nature of DM, and how to combine this information with all other searches. \\begin{figure*} \\centering \\includegraphics[width=\\textwidth]{dmmap2.eps} \\caption{2007 census of present and upcoming Dark Matter-related experiments. Black points denote the location of high energy neutrino telescopes; Dark-blue points are for gamma-ray Air Cherenkov Telescopes, while light-blue points are for other ground-based gamma-ray observatories. Red points are for underground laboratories hosting existing and upcoming direct detection experiments. Yellow points show the location of the Fermilab's Tevatron, and the upcoming Large Hadron Collider at CERN.} \\label{DMmap} \\end{figure*} ", "conclusions": "" }, "0710/0710.2215_arXiv.txt": { "abstract": "{Sequences of Doppler images of the young, rapidly rotating late-type stars AB Dor and LQ Hya show that their equatorial angular velocity and the amplitude of their surface differential rotation vary versus time. Such variations can be modelled to obtain information on the intensity of the azimuthal magnetic stresses within stellar convection zones. We introduce a simple model in the framework of the mean-field theory and discuss briefly the results of its application to those solar-like stars. } ", "introduction": "\\noindent Doppler imaging techniques allow us to measure the surface differential rotation in rapidly rotating late-type stars by tracking the longitudinal motion of starspots located at different latitudes (Collier Cameron 2007). Spe\\-ci\\-fi\\-cal\\-ly, the surface angular velocity $\\Omega$ at colatitude $\\theta$ is assumed to be given by a solar-like law: \\begin{equation} \\Omega(\\theta) = \\Omega_{\\rm eq} - d\\Omega \\cos^{2} \\theta, \\end{equation} where $\\Omega_{\\rm eq}$ is the equatorial angular velocity and $d\\Omega$ is the pole-equator angular velocity difference. $\\Omega_{\\rm eq}$ and $d\\Omega$ can be measured by fitting the shear of starspots along sequences of photospheric images covering successive rotations. Alternatively, $\\Omega$ and $d\\Omega$ can be included as free parameters in a code that reproduces the line profile distortions due to starspots, thus obtaining their values by a suitable $\\chi^{2}$ minimization when a sufficiently long time series of line profiles is available. The long-term monitoring of surface differential rotation of the two late-type dwarfs AB Doradus and LQ~Hydrae shows that their equatorial angular velocity and surface shear are functions of the time (see Donati et al. 2003a; Jeffers et al. 2007). It is interesting to note that the variations of $\\Omega_{\\rm eq}$ and $d\\Omega$ are compatible with an internal angular velocity uniform on cylindrical surfaces co-axial with the rotation axis (Donati et al. 2003a). ", "conclusions": "We modelled the observed time variation of the differential rotation in AB~Dor and LQ~Hya under the hypotheses that the azimuthal Maxwell stresses rule the changes of their surface shear and the internal angular velocity depends only on the distance from the rotation axis (Taylor-Proudman regime). We obtained that the average intensity of the mean field Maxwell stres\\-ses is $ |B_{s} B_{\\varphi} | \\sim 0.03 - 0.14 $ T$^{2}$, implying azimuthal mean fields $B_{\\varphi} \\sim (3-10)$ T for $B_{\\rm s} \\sim 0.01$ T. Similar Maxwell stresses are obtained if the magnetic torque is assumed to be confined within the overshoot layer $0.67 \\leq (r/R) \\leq 0.71$ and no restrictions are imposed on the internal rotation law. It is interesting to note that azimuthal magnetic fields of $3-10$ T, occupying a sizeable fraction of the convection zone, have been invoked to explain orbital period changes observed in late-type active binaries (Lanza, Rodon\\`o \\linebreak \\& Rosner 1998; Lanza \\& Rodon\\`o 2004; Lanza 2005, \\linebreak 2006b). An $\\alpha$-effect related to an instability of the magnetic field itself (e.g., magnetic buoyancy instability, Brandenburg \\& Schmitt 1998; or magneto-rotational instability, R\\\"udiger et al. 2007) seems to be necessary to produce such super-equi\\-par\\-ti\\-tion fields, possibly acting in combination with differential rotation. The energy dissipated by turbulence, estimated according to standard mixing-length arguments, may exceed stellar luminosity in the case of the largest surface shear observed in LQ Hya. However, the thermal equilibrium of the convection zone can be significantly affected only if those large shear episodes last more than $\\sim 10-20$ per cent of the time. Note also that a mixing-length estimate for the turbulent viscosity may not be appropriate for a rapidly rotating star (see Lanza 2006a for details). In the present work, we adopted a point of view analogous to that of Covas et al. (2005) who modelled the variation of stellar differential rotation considering only the \\linebreak torque exerted by the Lorentz force and neglecting the roles of meridional flow and $\\Lambda$-effect. As a matter of fact, it is difficult to evaluate the perturbations of the meridional flow and of the Reynolds stresses produced by the magnetic field because they depend critically on the approximations made in the treatment of stellar turbulence in a rotating star. Nevertheless, alternative models for the variation of differential rotation have been investigated, such as those based on a time-dependent component of the meridional flow (Rempel 2006, 2007) or the magnetic quenching of the $\\Lambda$-effect (R\\\"udiger et al. 1986). The present approach can be generalized to obtain amplitudes of the perturbations of the corresponding terms in Eq.~(\\ref{tau}), but we shall not pursue this application here (see Lanza 2007). Finally, it is interesting to note that some inference on the amplitude of variation of the surface shear in the case of very active stars can be obtained not only by means of Doppler imaging techniques, but also by an appropriate ana\\-ly\\-sis of their long-term wide-band photometry (e.g., \\linebreak Rodon\\`o et al. 2001; Messina \\& Guinan 2003). For example, in the case of LQ~Hya, it is worth comparing the Doppler imaging results by Donati et al. (2003a) with the photometrically determined seasonal rotation periods by Kov\\'ari et al. (2004)." }, "0710/0710.0210_arXiv.txt": { "abstract": "The unified dark energy and dark matter model within the framework of a model of a continuous medium with bulk viscosity (dark fluid) is considered. It is supposed that the bulk viscosity coefficient is an arbitrary function of the Hubble parameter. The choice of this function is carried out under the requirement to satisfy the observational data from recombination ($z\\approx 1000$) till present time. ", "introduction": "Modelling of an accelerated expansion of the present Universe lies on the way of creation of phenomenological models which may explain the observational data on one set of parameters and compare them with predictions of the models on other set. For example, a theoretical model adapts for the correct description of the acceleration of the Universe in an accessible interval $z$. Further, the results of modelling are extrapolated for large $z$ which are not accessible for observations yet. The corresponding cosmological scenario defines growth of the large-scale structure which determines the present-day fluctuations of the microwave background radiation. Certainly, the models under consideration should not contradict the available observational data within the framework of general relativity in a field of its applicability. For the specified purposes a number of cosmological models successfully applied in the past in the theory of the early Universe is used. These are cosmological models with various scalar and non-scalar fields filling the space together with cold dark matter (see, e.g., the reviews \\cite{Sahni}). Cosmological models of the present accelerated Universe within the framework of high-order theories of gravity (HOTG) are also quite popular \\cite{Carrol}. A number of models have recently been suggested \\cite{Ren, Brevik,Odin} which describe the present Universe with use of models of a continuous medium in the presence of bulk viscosity. Consideration of effects of viscosity within the framework of HOTG was also carried out \\cite{Brevik1}. Note that such models were well-known in the theory of the early Universe (see, for example, \\cite{Murphy,Barrow}). In particular, in Ref. \\cite{Barrow} a few exact solutions with the constant bulk viscosity coefficient and with the bulk viscosity being an arbitrary power function of energy density were obtained. In Ref. \\cite{Ren} the model of viscous dark fluid is considered. The main result of this paper is the model with the constant bulk viscosity coefficient. The model fits the observational data on luminosity at an acceptable level. In Ref. \\cite{Brevik} the models both with the constant bulk viscosity coefficient and the bulk viscosity linearly proportional to the Hubble parameter are examined. The question about influence of viscosity on presence of a singularity in the model in the future (the so-called Big Rip) is investigated. In this paper we consider a model of \"viscous dark fluid\" with the bulk viscosity coefficient $\\mu(H)$ which depends on the Hubble parameter arbitrarily. Unlike Ref. \\cite{Ren}, comparison of the model with the observational data is not restricted to the observational data on luminosity. The model is being compared with results of observations on change of the deceleration parameter $q$ and values of the Hubble parameter in the range $2>z>0$. It will be shown below that the model with the constant bulk viscosity coefficient does not provide a good description for $q(z)$ and $H(z)$ which follow from the observations. We propose such a dependence $\\mu(H)$ which is adequate to the mentioned observations. The proposed model is extrapolated for $z$ beyond the specified range $2>z>0$. ", "conclusions": "The bulk viscosity in our model is an example of a dynamic $\\Lambda$-term. However, our model is close to the $\\Lambda$CDM model. As it was rightly noted in \\cite{Ren}, the model with the viscosity does not give possibility to divide the true dust filling the Universe, and dark matter generated by the bulk viscosity. That is why it is difficult to introduce a phenomenological equation of state $p=w \\varepsilon$ which is often used for interpretation of the observational data. Influence of the bulk viscosity on formation of the large-scale structure of the Universe demands special examination. But taking into account that the viscosity is being ''involved`` at rather small $z\\approx 2$, it is possible to expect its influence on the dynamics of galactic clusters only at later non-linear stage." }, "0710/0710.5435_arXiv.txt": { "abstract": "{\\object{NRAO~150} --a compact and bright radio to mm source showing core/jet structure-- has been recently identified as a quasar at redshift $z=1.52$ through a near-IR spectral observation.} {To study the jet kinematics on the smallest accessible scales and to compute the first estimates of its basic physical properties,} {we have analysed the ultra-high-resolution images from a new monitoring program at 86~GHz and 43~GHz with the GMVA and the VLBA, respectively. An additional archival and calibration VLBA data set, covering from 1997 to 2007, has been used.} {Our data shows an extreme projected counter-clock-wise jet position angle swing at an angular rate of up to $\\approx 11^{\\circ}/\\rm{yr}$ within the inner $\\approx 31$~pc of the jet, which is associated with a non-ballistic superluminal motion of the jet within this region.} {The results suggest that the magnetic field could play an important role in the dynamics of the jet in \\object{NRAO~150}, which is supported by the large values of the magnetic field strength obtained from our first estimates. The extreme characteristics of the jet swing make \\object{NRAO~150} a prime source to study the jet wobbling phenomenon.} ", "introduction": "\\label{int} An increasing number of jets in active galactic nuclei (AGN) have been reported to show either regular or irregular swings of the innermost jet structural position angle in the plane of the sky (e.g., in \\object{OJ~287}, Tateyama \\& Kingham~\\cite{Tat04}; in \\object{3C~273}, Savolainen et al.~\\cite{Sav06}; in \\object{3C~345}, Lobanov \\& Roland~\\cite{Lob05}; in \\object{BL~Lac}, Stirling et al.~\\cite{Sti03}; in \\object{S5~0716+71}, Bach et al.~\\cite{Bac05}). Time scales between 2 and 15 years and structural position-angle oscillations with amplitudes from $\\sim 25^{\\circ}$ to $\\sim 45^{\\circ}$ are typical for the reported cases. We will call this phenomenon \\emph{jet wobbling} hereafter. Parsec scale AGN jet curvatures and helical-like structures also at larger distances from the central engine are also believed to be triggered by changes in direction at the jet ejection nozzle (e.g., in \\object{3C~84}, Dhawan et al.~\\cite{Dha98}). The physical origin for the observed jet wobbling is still poorly understood. Among the various possibilities, regular precession of the accretion disk is frequently used for modeling at present. Most AGN precession models are driven either by a companion super-massive black hole or another massive object inducing torques in the accretion disk of the primary (e.g., Lister et al.~\\cite{Lis03}, for \\object{4C~+12.50}; Stirling et al.~\\cite{Sti03}, for \\object{BL~Lac}; Caproni \\& Abraham~\\cite{CapAb04} for \\object{3C~120}) or by the Bardeen-Peterson effect (e.g., Liu \\& Melia~\\cite{Liu02}; Caproni et al.~\\cite{Cap04}). However, other AGN scenarios that have yet to be explored extensively, such as the orbital motion of the jet nozzles (also involving binary systems) or other kinds of more erratic disk/jet instabilities (e.g., similar to those thought to produce the quasi periodic oscillations [QPO] in X-ray binaries), can not be ruled out yet. Note that, in support of these erratic instabilities, it is still under debate whether the observed jet wobbling is strictly periodic or not (see Mutel \\& Denn~\\cite{Mut05} for the case of \\object{BL~Lac}). There is still no paradigm to explain the phenomenon of jet wobbling in AGN, but it is rather likely that, as it is triggered in the innermost regions of the jets, it must be tied to fundamental properties of the inner regions of the accretion system. Hence, there is ample motivation to study the jet wobbling phenomenon to place our understanding of the jet triggering region and the super--massive accretion systems on firmer ground. VLBI observations at millimetre wavelengths are a powerful technique to image the innermost regions of AGN jets -which are self-absorbed at longer wavelengths- with the highest angular resolutions; $\\sim 50\\,\\mu$as at 86~GHz (3.5~mm) and $\\sim 0.15$\\,mas at 43~GHz (7~mm). Here, we report the discovery of an extreme case of jet swing in the quasar \\object{NRAO~150}, through the first ultra-high-resolution VLBI set of images obtained from this source at 86~GHz and 43~GHz. \\object{NRAO~150} is a strong radio-mm source. At radio wavelengths, on VLBI scales, \\object{NRAO~150} displays a compact core plus a one-sided jet extending up to $r \\apgt 80$~mas with a jet structural position angle (PA) of $\\sim 30^\\circ$ (e.g., Fey \\& Charlot~\\cite{Fey00}). \\object{NRAO~150} was not detected by the early optical surveys, most probably due to obscuration through the Milky Way (Galactic latitude $= -1.6^\\circ$). Almost 40~yr after its discovery at radio wavelengths (Pauliny-Toth, Wade \\& Heeschen~\\cite{Pau66}), \\object{NRAO~150} has been identified as a quasar at redshift $z=1.52$ through a near-IR spectroscopic project (Acosta-Pulido et al. in prep.). Throughout this paper we assume $H_{\\circ} = 72$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{m}=0.3$, and $\\Omega_{\\Lambda}=0.7$. Under these assumptions, the luminosity distance of \\object{NRAO~150} is $d _{L}=11025$~Mpc, 1~mas corresponds to $8.5$~pc in the frame of the source, and an angular proper motion of 1~mas/yr translates into a speed of $69.4~c$. These parameters are used here for the first time to make quantitative estimates of the basic physical properties of the jet in \\object{NRAO~150}. ", "conclusions": "\\label{concl} We have reported the results from the first 86~GHz and 43~GHz VLBI monitoring program of the recently identified quasar \\object{NRAO~150}. Our observations reveal {\\it a)} a large projected misalignment of the jet by $>100^{\\circ}$ within the inner 0.5~mas to 1~mas from the core, {\\it b)} an extremely fast counter-clockwise rotation of the projected jet axis at a rate of $\\sim6^{\\circ}/\\rm{yr}$ to $\\sim11^{\\circ}/\\rm{yr}$, {\\it c)} non-ballistic superluminal motions with mean speeds from 2.3~$c$ to 3.3~$c$ within the inner 0.5~mas from the core (deprojected distance $\\approx 31$~pc), {\\it d)} transverse (non-radial) speeds of 2.7~$c$, 1.4~$c$, and 2.0~$c$ for Q1, Q2 and Q3, respectively, {\\it e)} an extreme case of \\emph{superluminal non-ballistic jet swing}, and {\\it f)} the first approximations to quantitative estimates of the basic physical properties of the jet in \\object{NRAO~150}: $\\delta \\approx 6$, $B\\approx 0.7$\\,G, $\\gamma \\approx 4$, and $\\phi \\approx 8^{\\circ}$. Whereas the ultimate origin of the jet swing must be an intrinsic change of the direction of the inner jet axis (either caused by changes produced at the injection region or by interaction with the medium surrounding the jet), possible causes for the non-ballistic nature of the emitting flow are either a small inertia of the jet compared to that of the impacted ambient medium, or by the fact that we are observing a jet instability propagating downstream. However, the superluminal non-ballistic motion of the jet features in \\object{NRAO~150} is perhaps too fast and systematic along tens of parsecs (deprojected, see above) to be induced solely by the ambient medium. In addition, the perturbation that must be produced either by the impact with such medium or by the changes at the injection region should imply the growth of disruptive instabilities (e.g., Perucho et al.~\\cite{Per06}), in contrast to the remarkable collimation of the jet up to the kiloparsec scale (see Fig.~1). Mizuno et al.~(\\cite{Miz07}) have shown that a magnetic field with the appropriate configuration helps keeping a jet collimated against the growth of instabilities. Hence, this possibility, together with the high value for the magnetic field intensity estimated in Section~\\ref{phys-par}, suggests that the magnetic field could play an important role in the dynamics of the jet in \\object{NRAO~150}. It is still unclear whether the reported change of the direction of ejection in \\object{NRAO~150} is related to a regular (strictly periodic or not) behaviour or to a single event. This, together with the still unknown nature of the moving jet features (either propagating curvatures or shocks in a bent jet), does not allow us to specify the ultimate origin of this phenomenon in the source. This will be the matter of future studies. Nevertheless, the extreme characteristics of the jet swing make \\object{NRAO~150} a prime source for future studies of the origin of jet wobbling." }, "0710/0710.0817_arXiv.txt": { "abstract": "The fast rotating star CU Virginis is a magnetic chemically peculiar star with an oblique dipolar magnetic field. The continuum radio emission has been interpreted as gyrosyncrotron emission arising from a thin magnetospheric layer. Previous radio observations at 1.4~GHz showed that a 100\\% circular polarized and highly directive emission component overlaps to the continuum emission two times per rotation, when the magnetic axis lies in the plane of the sky. This sort of radio lighthouse has been proposed to be due to cyclotron maser emission generated above the magnetic pole and propagating perpendicularly to the magnetic axis. Observations carried out with the Australia Telescope Compact Array at 1.4 and 2.5~GHz one year after this discovery show that this radio emission is still present, meaning that the phenomenon responsible for this process is steady on a timescale of years. The emitted radiation spans at least 1 GHz, being observed from 1.4 to 2.5~GHz. On the light of recent results on the physics of the magnetosphere of this star, the possibility of plasma radiation is ruled out. The characteristics of this radio lighthouse provides us a good marker of the rotation period, since the peaks are visible at particular rotational phases. After one year, they show a delay of about 15 minutes. This is interpreted as a new abrupt spinning down of the star. Among several possibilities, a quick emptying of the equatorial magnetic belt after reaching the maximum density can account for the magnitude of the breaking. The study of the coherent emission in stars like CU~Vir, as well as in pre main sequence stars, can give important insight into the angular momentum evolution in young stars. This is a promising field of investigation that high sensitivity radio interferometers such as SKA can exploit. ", "introduction": "CU Virginis (=HD124224) is an A-type magnetic chemically peculiar star (MCP) with a rotational period of 0.52 days, one of the shortest for this class of stars. As observed in all MCP stars, the variability of the light curve is correlated with the spectroscopic variations \\citep{deutsch, hardie} and with the effective magnetic field \\citep{borra80}. All those characteristics can be explained in the framework of the oblique rotator model, where the axis of the dipolar magnetic field is tilted with respect to the rotational one \\citep{babcock} and abundance of elements is not homogeneously distributed over the stellar surface. The observed variabilities are thus consequence of the stellar rotation. \\begin{figure*} \\includegraphics[width=17cm]{spettri.ps} \\caption{Dynamical spectra of CU~Vir during the two days of observations (left panels: May 29, 1999) at 2.5~GHz (upper panels) and 1.4~GHz (lover panels). Some spectral channels have been removed, reducing the bandpass. The spectral resolution is 8~MHz; the spectra have been smoothed in time with a window 2 minutes wide. Strong enhancements of the radio emission are evident at two phases at 1.4~GHz, while only one peak is visible at 2.5~GHz. } \\label{spettri} \\end{figure*} MCP stars are in general slow rotators in comparison with normal B and A-type main sequence stars. This behaviour can be explained as the result of the action of a magnetic breaking. But, at the present, only two MCP stars have been found to increase their rotational period; they are 56 Ari and CU Vir. While the former shows a continuous breaking down at the rate of few seconds per century \\citep{adelman}, CU Vir has been subject of an abrupt increase of the rotational period \\citep{pyper}. From the analysis of photometric light curves collected over 40 years, it seems that a change of period of about two seconds occurred abruptly in 1984. The mechanism responsible for this event is still under debate, and precise timing of the rotation is needed in order to detect any further slowing down of the star. Radio emission has been observed in CU Vir \\citep{leo94}. The radio spectrum is quite flat and extends up to mm wavelengths \\citep{leo2004} with an high degree of circular polarization \\citep{leo96}. The variabilities of total intensity and polarization are both correlated with the rotation of the star, suggesting we are in presence of gyrosyncrotron emission from an optically thick source. The anisotropic stellar wind inferred from spectral line variations in MCP star \\citep{shore}, associated with the magnetic field justifies the radio emission. \\citet{trig04} developed a three-dimensional model to explain the radio emission from MCP stars. The dipolar magnetic field interacts with the stellar wind, that can freely escape from the polar regions outside the Alfv\\'en surface (outer magnetosphere) but remains trapped in the equatorial belt (inner magnetosphere). Ionized particles flowing out in the transition region between outer and inner magnetosphere, called middle magnetosphere, stretch the magnetic field lines just outside the Alfv\\'en radius and open the field generating current sheets, where particles are accelerated up to relativistic energies. They eventually propagate back toward the stellar magnetic poles following the field lines and radiating for gyrosyncrotron process. As the magnetic field intensity increases traveling to the star, they are reflected back outward by the magnetic mirroring effect and are definitively lost from the magnetosphere. This model, used to explain the observed fluxes and variability of MCP stars (HD37479 and HD37017) has been successfully applied to CU Vir by \\citet{leto06}. On the basis of multiwavelengths radio observations, important physical parameters of the stellar magnetosphere, as the Alfv\\'en radius ($12-17\\,R_\\ast$) and the mass loss (about $10^{-12}M_\\odot \\rmn{yr}^{-1}$) have been inferred. A further observational evidence supporting the picture outlined above is the discovery of coherent, highly directive, 100\\% polarized radio emission at 1.4~GHz \\citep{trig00}. The two peaks of radio emission have been observed at the rotational phases when the magnetic axis of the dipole is perpendicular to the line of sight. The two peaks have been observed in three observing runs spanning 10 days, indicating that the emission mechanism is persistent at least in timescales of weeks. This has been interpreted as Electron Cyclotron Maser Emission from a population of electrons accelerated in the current sheets at the Alfv\\'en point, that developed a loss cone anisotropy after the mirroring and masing in a direction almost perpendicular to the motion, so to the field lines, just above the magnetic pole. In this paper we present radio observations of CU Vir at 1.4 and 2.5~GHz carried out with the aim to confirm the maser emission and to study its spectrum. The directivity of the radiation is used to check the rotational period as the beam point toward the Earth two times per rotation. ", "conclusions": "The observations of CU Vir carried out with the ATCA at 1.4 and 2.5 GHz confirm the presence of the coherent emission already reported by \\citet{trig00} after one year from the discovery, indicating that this is a steady phenomenon. The emitted radiation is visible only at rotational phases corresponding to the instant when the oblique axis of the magnetic dipole lies on the plane of the sky. This indicates the high directivity of this component of the radio emission. All those characteristics, and the fact that it is 100\\% right hand circularly polarized, are in agreement with the process of electron cyclotron maser from electrons accelerated in the current sheets out of the Alfv\\'en radius, propagating back to the stellar polar caps and reflected outward by the converging magnetic field. The lack of reflected electrons at small pitch angle, that fall in the stellar atmosphere, is the cause of the loss cone anisotropy that, in turn, generates this auroral radiation above the magnetic pole of the star. The polarization properties, the fact that the radiation is only right hand polarized, means that this process in efficient only above the north magnetic pole. This is possible as the magnetic field is not purely dipolar. The possibility of plasma radiation is ruled out since the high magnetic field strength in the region where the radiation is generated. Since the magnetosphere rotates obliquely around the rotational axis, the cyclotron maser is visible only when it points toward the Earth, like a lighthouse. In this mode it is a good marker of the rotation of the star. The analysis done shows that the peaks are delayed of about 15 minutes with respect to the observations carried out one year before, indicating a possible change of the rotational period of the star, of the order of 1 second, occurred in the period 1998--1999. A similar change of period \\citep{pyper}, occurred around 1985, has been already reported. CU Vir is the unique single main sequence star with frequent abrupt spindown. Different spinning down mechanisms are discussed: the possibility of a change of the moment of inertia of the star, the continuous spindown due to the wind flowing from the Alfv\\'en surface and the violent emptying of the inner magnetosphere. In this latter hypothesis the material accumulated in the closed field lines of the equatorial magnetic belt reaches a maximum density and opens the field lines in a violent event, releasing an angular momentum which may account the observed breaking. Cyclotron maser emission from stars provides important information on the magnetospheres, as it has been observed in flare stars, in dMe and brown dwarfs. In the future, when high sensitivity radio interferometers such as SKA will allow to discover more and more radio lighthouse of the same kind of CU Vir, a new possibility to study with high precision the angular momentum evolution of young main sequence and pre main sequence stars will be opened." }, "0710/0710.1811_arXiv.txt": { "abstract": "{}{X-ray Bright Optically Normal Galaxies (\\xbongs) constitute a small but not negligible fraction of hard X-ray selected sources in recent \\chandra\\ and \\xmm\\ surveys. Even though several possibilities were proposed to explain why a relatively luminous hard X-ray source does not leave any significant signature of its presence in terms of optical emission lines, the nature of \\xbongs\\ is still subject of debate. We aim to a better understanding of their nature by means of a multiwavelength and morphological analysis of a small sample of these sources. } {Good-quality photometric near-infrared data (ISAAC/VLT) of four low-redshift ($z=0.1-0.3$) \\xbongs, selected from the {\\rm HELLAS2XMM} survey, have been used to search for the presence of the putative nucleus, applying the surface-brightness decomposition technique through the least-squares fitting program GALFIT. } {The surface brightness decomposition allows us to reveal a nuclear point-like source, likely to be responsible of the X-ray emission, in two out of the four sources. The results indicate that moderate amounts of gas and dust, covering a large solid angle (possibly 4$\\pi$) at the nuclear source, combined with the low nuclear activity, may explain the lack of optical emission lines. The third \\xbong\\ is associated with an X-ray extended source and no nuclear excess is detected in the near infrared at the limits of our observations. The last source is associated to a close (d$\\leq$ 1 arcsec) double system and the fitting procedure cannot achieve a firm conclusion.}{} ", "introduction": "Thanks to the \\chandra\\ and \\xmm\\ surveys, the hard X-ray sky is now probed down to a flux limit where the bulk of the X-ray background is almost completely resolved into discrete sources (Hasinger et al. 2001; Alexander et al. 2003; Bauer et al. 2004; Worsley et al. 2004, 2005). Extensive programs of multiwavelength follow-up observations showed that the large majority of hard X-ray selected sources are identified with Active Galactic Nuclei (AGN) spanning a broad range of redshifts and luminosities. At variance with optically selected quasars, X-ray selected AGN are characterized by a much larger spread in their optical properties, especially for what concerns the intensity of the emission lines. Indeed, a sizable fraction of relatively luminous X-ray sources hosting an active nucleus would not have been easily recognized as such on the basis of optical observations either because associated with very faint ($R >$ 24) counterparts (e.g., Fiore et al. 2003; Mignoli et al. 2004; Civano et al. 2005) or due to the lack of AGN emission lines in their optical spectra. The latter class of sources is variously termed as ``optically-dull\", ``optically normal\" or \\xbongs\\ (X-ray Bright Optically Normal Galaxies; Comastri et al. 2002). The common meaning of these definitions is that they lack evidence of accretion-driven activity in their optical spectra, in contrast with ``normal\" Seyfert galaxies and quasars. Their X-ray luminosities ($\\approx 10^{42}-10^{43}$ \\lum), X-ray spectral shape and X-ray-to-optical flux ratio (X/O\\footnote{Where X/O is defined as $X/O=\\log{\\frac{f_X}{f_R}}=\\log{f_X}+C+\\frac{R}{2.5}$.}$\\sim -1$) suggest AGN activity of moderate strength. Originally discovered in early {\\it Einstein} observations (Elvis et al. 1981) and named optically dull galaxies, the interest on the nature of these sources has gained a renewed attention after the discovery of several examples in XMM-{\\it Newton} and {\\it Chandra} surveys (Fiore et al. 2000; Comastri et al. 2002a,b; Georgantopoulos et al. 2005; Kim et al. 2006). Several possibilities were proposed in the literature in order to explain why a relatively luminous, hard X-ray source does not leave any significant signature of its presence in the form of emission lines. A simple explanation favoured by Moran et al. (2002) and more recently by Caccianiga et al. (2007) for faint sources in the \\chandra\\ deep fields and brighter object in the \\xmm\\ XBS survey, respectively, is dilution by the host galaxy starlight. The combination of optical faintness and lack of strong emission lines in the observed wavelength range for distant \\chandra\\ sources or the inadequate observing set-up among brighter nearby \\xmm\\ objects (Severgnini et al. 2003) may account for the \\xbong\\ properties. More in general, if the contrast between the host galaxy starlight and nuclear emission is high, AGN emission lines may easily be undetected. The physical reason may be ascribed to obscuration, most likely with a large covering factor, or to the fact that the lines are not efficiently produced by the central engine. If \\xbongs\\ are merely obscured AGN, two hypotheses may be envisaged: \\par $\\bullet$ In order to explain the multiwavelength properties of the \\xbong\\ prototype PKS~0312018, also known as P3, Comastri et al. (2002) suggested heavy obscuration by Compton-thick gas covering almost 4$\\pi$ at the nuclear X-ray sources. In this way, no ionizing photons can escape to produce the narrow emission lines which are observed in ``normal\" Type 2 narrow-line AGN which are thought to have a lower covering fraction following the AGN Unified Scheme (but see Section \\ref{xray} for a recent re-analysis of the X-ray data of P3). \\par $\\bullet$ According to a detailed multiwavelength analysis of ``optically-dull\" galaxies in the \\chandra\\ deep fields, Rigby et al. (2006) conclude that extranuclear dust in the host galaxy plays an important role in hiding the emission lines.\\\\ Alternatively, \\xbongs\\ may be members of a class, or classes, of {\\it exotic} objects for which emission lines are either intrisically weak or absent: \\par $\\bullet$ Radiatively Inefficient Accretion Flows (RIAFs) are expected at accretion rates well below those inferred for Seyferts and quasars. A distinctive property of low accretion-rate flows is that the standard Shakura-Sunyaev accretion disk is truncated at a relatively large inner radius. As a consequence, it cannot generate the ``big blue bump\" and enough UV photons to photoionize the line-emitting circumnuclear gas. The infalling gas is heated to high temperatures and emits a hard X-ray power-law by upscattering of low-energy seed photons. According to Yuan \\& Narayan (2004), the SED of source P3 could be reproduced by a RIAF model. \\par $\\bullet$ \\xbongs\\ could be extreme BL Lac objects in which the featureless non-thermal continuum is much weaker than the host galaxy starlight. Following Fossati at al. (1998), \\xbongs\\ could belong to the low-luminosity tail of the blazar spectral sequence based on the anti-correlation between luminosity and frequency of the synchrotron peak. \\par $\\bullet$ A highly speculative possibility is that of a transient AGN phenomenon in the process of tidally disrupting a star. If this were the case, the X-ray emission should be witnessing the transient accretion phenomenon (see Komossa et al. 2004 for extreme variability events in ROSAT observations; see also Gezari et al. 2006 for a luminous flare observed in the GALEX Deep Imaging Survey). The transient is most likely over in subsequent follow-up optical observations. Finally, it should be noted that diffuse emission from a galaxy group, whose X-ray extended emission may have escaped detection in low signal-to-noise X-ray observations, is also possible and indeed observed in a few cases (Georgantopoulos et al. 2005). While a unique solution may not necessarily hold for all the \\xbongs\\ observed in different surveys, they represent a useful benchmark for a better understanding of the AGN activity and, as such, deserve further studies. Ideally, one would need sensitive, high-spatial resolution, multiwavelength observations from radio to X-rays. As a first step, in the following we use good-quality photometric near-infrared data obtained with ISAAC at VLT of four low-redshift ($z=0.1-0.3$) \\xbongs, selected in the {\\rm HELLAS2XMM} survey to search for the presence of a putative nucleus which has escaped detection in the optical spectroscopy. The rather obvious advantage of near-infrared data is that the effects of dust reddening are minimized and that the nuclear emission in this band should rise more rapidly than the stellar light due to the reprocessing by hot dust. At the same time, the excellent quality of the near-IR images makes possible to apply a surface brightness decomposition technique, already successfully applied by several authors (e.g. S\\'anchez et al. 2004; Peng et al. 2006), to search for weak unresolved nuclear emission down to faint near-infrared magnitudes. We also discuss the broad-band properties of the four \\xbongs\\ using available multiwavelength data. Throughout the paper we assume a cosmological model with $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_m$ = 0.3 and $\\Omega_{\\Lambda}= 0.7$. ", "conclusions": "We have presented a multiwavelength analysis of four \\xbongs\\ selected from the {\\rm HELLAS2XMM} survey. For these sources, deep near-infrared images taken with ISAAC at VLT, good-quality optical spectra and \\xmm\\ data are available. Applying the morphological decomposition technique, we were able to detect the presence of a nuclear component in two out of the four sources (PKS~03120017 and PKS~03120018). There is no evidence of nuclear emission in the near-infrared in Abell~2690013; moreover, the X-ray appearance is consistent with an extended source. For source Abell~1835140, the near-infrared images reveal a complex morphology, where two sources are embedded in a common envelope. The main issue about the \\xbong\\ nature is whether they represent a truly distinct class or, rather, they are a mixed source population. Our results point towards the latter hypothesis. The results regarding the nature of the \\xbongs\\ with nuclear component and the lack of optical emission lines can be summarized as follows. \\begin{itemize} \\item Source PKS~03120017 and PKS~03120018 are well described by a mildly obscured ($E(B-V)= 0.5-0.8$) optically weak nucleus, responsible for the X-ray emission, hosted by a bright galaxy (mag$_{K_s}^{nucl}$--mag$_{K_s}^{host}$ $\\sim 4$). \\item The lack of optical emission lines cannot be attributed to observational limitations such as a inadequate observational setup or low signal-to-noise-ratio, at least for the objects in the present sample for which high-quality spectroscopic and photometric observations are available. \\item We can safely discard for the two objects both a Compton-thick scenario (as found by Caccianiga et al. 2007 for a sample of elusive AGN in the XBSS) and an important blazar contribution. A RIAF solution seems to be supported by the estimated values of the Eddington ratios in the two objects and is not ruled out by the broad-band SED fitting, also because of the large number of free parameters in the RIAF model that can be tuned in order to reproduce the observed SEDs. A weak nuclear source, described by a standard accretion disk solution, but not powerful enough in the production of UV photons, would also provide an acceptable description of the observations. \\item The presence of a thin nuclear gas and dust structure (as argued by Cocchia et al. 2007) covering 4$\\pi$ at the nuclear source, combined with the low level of activity of the BH, could prevent the ionization of the narrow-line regions and produce also the extinction we measured.\\\\ The analysis of already obtained VLT/VIMOS-IFU (Integral Field Unit) observations could help in the detection of the typical AGN emission lines down to a very faint flux limit and in the study of the gas kinematics. \\end{itemize}" }, "0710/0710.0671.txt": { "abstract": "Astronomy provides a laboratory for extreme physics, a window into environments at extremes of distance, temperature and density that often can't be reproduced in Earth laboratories, or at least not right away. A surprising amount of the science we understand today started out as solutions to problems in astronomy. Some of this science was key in the development of many technologies which we enjoy today. This paper describes some of these connections between astronomy and technology and their history. ", "introduction": " ", "conclusions": "" }, "0710/0710.3163_arXiv.txt": { "abstract": "This template file shows how to use the \\texttt{aipproc} class to produce a paper with the correct layout for \\emph{% AIP Conference Proceedings 8.5in x 11in double column}. A full description of the features supported by the \\texttt{aipproc} class can be found in the \\texttt{aipguide.pdf} document accompanying the distribution. Frequently asked questions can be found in the \\texttt{FAQ.txt} document. ", "introduction": "Infandum, regina, iubes renovare dolorem, Troianas ut opes et lamentabile regnum cruerint Danai; quaeque ipse miserrima vidi, et quorum pars magna fui. Quis talia fando Myrmidonum Dolopumve aut duri miles Ulixi temperet a lacrimis? Et iam nox umida caelo praecipitat, suadentque cadentia sidera somnos. ", "conclusions": "" }, "0710/0710.5486_arXiv.txt": { "abstract": "{ Type~Ia supernovae are believed to be white dwarfs disrupted by a thermonuclear explosion. Here we investigate the scenario in which a rather low-mass, carbon-oxygen (C~+~O) white dwarf accumulates helium on its surface in a sufficient amount for igniting a detonation in the helium shell before the Chandrasekhar mass is reached. In principle, this can happen on white dwarfs accreting from a non-degenerate companion or by merging a C~+~O white dwarf with a low-mass helium one. In this scenario, the helium detonation is thought to trigger a secondary detonation in the C~+~O core. It is therefore called the ``double-detonation sub-Chandrasekhar'' supernova model. By means of a set of numerical simulations, we investigate the robustness of this explosion mechanism for generic 1-\\msol\\ models and analyze its observable predictions. Also a resolution dependence in numerical simulations is analyzed. Hydrodynamic simulations of the double-detonation sub-Chandrasekhar scenario are conducted in two and three spatial dimensions. The propagation of thermonuclear detonation fronts, both in helium and in the carbon-oxygen mixture, is computed by means of both a level-set function and a simplified description for nuclear reactions. The decision whether a secondary detonation is triggered in the white dwarf's core or not is made based on criteria given in the literature. In a parameter study involving different initial flame geometries for He-shell masses of 0.2 and 0.1~\\msol\\ (and thus 0.8 and 0.9~\\msol\\ of C~+~O), we find that a secondary detonation ignition is a very robust process. Converging shock waves originating from the detonation in the He shell generate the conditions for a detonation near the center of the white dwarf in most of the cases considered. Finally, we follow the complete evolution of three selected models with 0.2~\\msol\\ of He through the C/O-detonation phase and obtain \\nni-masses of about 0.40 to 0.45~\\msol. Although we have not done a complete scan of the possible parameter space, our results show that sub-Chandrasekhar models are not good candidates for normal or sub-luminous type~Ia supernovae. The chemical composition of the ejecta features significant amounts of \\nni\\ in the outer layers at high expansion velocities, which is inconsistent with near-maximum spectra.} ", "introduction": "\\label{sec:int} One of the major uncertainties in modeling type~Ia supernovae (SN~Ia) originates from the unknown nature of the progenitor systems because neither observations nor theoretical models are yet conclusive \\citep{Branch1995,Ruiz-Lapuente2000,Livio2000,Nomoto2003,Han2004,Napiwotzki2005,Stritzinger2006,Parthasarathy2007}. Over the past years, most studies of thermonuclear supernova explosions focused on models in which a thermonuclear flame is formed by a runaway near the center of a white dwarf (WD), composed of carbon and oxygen, once it has reached the Chandrasekhar mass ($M_\\mathrm{Ch} \\sim 1.4~\\msol$) by accretion from a non-degenerate companion. Starting out in the sub-sonic deflagration mode of flame propagation, these models were shown to give rise to energetic explosions either in pure turbulence-boosted deflagrations \\citep{Reinecke2002b,Gamezo2003,Roepke2005,Roepke2006a,Schmidt2006a,Schmidt2006b,Roepke2007c} or with a delayed triggering of a supersonic detonation phase \\citep{Gamezo2004,Plewa2004,Roepke2007a}. There seems to be reasonable agreement of such models with the gross features of observed SNe~Ia \\citep{Roepke2007c,Mazzali2007}. However, it remains unclear whether the Chandrasekhar mass model can account for the full range of SN~Ia observations. In particular, the mechanism of the sub-luminous events (\\object{SN~1991bg} being a prototypical example) remains a puzzle \\citep{Mazzali2007}. \\citet{Stritzinger2006} even claim typical ejecta masses below $M_\\mathrm{Ch}$ for a sample of well observed SNe~Ia, with the trend that the low-luminosity explosions eject less mass. Therefore, a long standing question is whether other progenitor channels contribute to the observed SN~Ia sample, and two alternatives have been suggested: the \\emph{WD merger} or \\emph{double degenerate scenario} (in contrast to single degenerate models, which consist of binaries with only one WD) and the \\emph{sub-Chandrasekhar model}. In the present work we explore the latter. The basic idea of the sub-Chandrasekhar model is that if the accretion rate onto a WD is lower than about $1$ - $4 \\cdot 10^{-8}~\\msol \\, \\mathrm{yr}^{-1}$ the accreted He (or the H processed to He) is not fused steadily into C and O. Instead, after reaching a critical amount of He at relatively low densities, the He shell becomes unstable and detonates \\citep[cf.][and references therein]{Woosley1986}. The detonation ignites most likely close to the bottom of the He shell, produces almost pure \\nni, and can occur long before the WD reaches the Chandrasekhar limit. A second detonation may be triggered spontaneously when the He shell detonation shock wave hits the C~+~O core or with some delay after the shock has converged near the center of the WD (\\emph{double detonation} models). This way, in the sub-Chandrasekhar mass models the conditions for thermonuclear runaway are caused by the compression in the shock wave and not by the high degree of degeneracy as in the Chandrasekhar mass models. Delayed double detonations have been studied extensively before. \\citet{Woosley1994}, and \\citet{Livne1990} carried out one-dimensional (1D) simulations and \\citet{Livne1990a,Livne1991} considered two-dimensional (2D) setups. In 1D the He detonation ignites synchronously in a layer close to the bottom of the He shell and due to the spherical symmetry constraint the shock converges perfectly in the center. However, this is not a very realistic model. According to \\citet{Livne1990} an ignition most likely happens in a single point leading to an off-center convergence of oblique shock waves in the core. Several other simulations predict successful directly ignited double detonations in two dimensions \\citep{Livne1997,Arnett1997,Wiggins1997,Wiggins1998}. The smoothed particle hydrodynamics simulations by \\citet{Benz1997} and \\citet{Garcia-Senz1999} were carried out in 3D\\@. \\citet{Livne1990a} first reported an increased probability of the direct core ignition if the He detonation does not happen directly at the core--shell interface but at a certain distance above it. This way, the pressure jump can grow large enough before hitting the core. In this work, the results of a (restricted) parameter study are presented investigating the possibility of triggering the second detonation in two and three dimensions. In all our models the WD has a total mass of 1~\\msol, but the (C~+~O)-core and the He-shell masses differ. We compute sequences of models with very different initial flame geometries in order to test the robustness of this explosion mechanism. Since this latter question is the main focus of our paper, most of the simulations are stopped once the conditions for a detonation are matched. However, for a few successful cases the energetics and nucleosynthesis of complete double detonations are computed. In the following chapter the model details will be described. Section~\\ref{sec:sim} shows the simulation results, which are summarized and discussed in the last part of the paper. ", "conclusions": "\\label{sec:con} We have studied the sub-Chandrasekhar model of SNe~Ia by means of a series of two-dimensional and a few three-dimensional simulations with different initial conditions. The numerical scheme used is based on the PPM algorithm, and the propagation of the detonation front is modeled applying the level set technique. This novel implementation allowed for an accurate treatment of the hydrodynamic features of the sub-Chandrasekhar model (such as shock waves and thin detonation fronts) in multiple dimensions. With our generic 1-\\msol\\ model we performed a parameter study involving different He masses and ignition geometries of the initial He detonation. We find that a detonation in the He shell can clearly trigger a second detonation in the core and at least for the mass configurations studied, the double detonation seems to be a very robust process, which works without any ``fine-tuning'' of our model parameters. In almost all of the simulations performed, the ignition conditions for a core detonation were reached at or near the center of the WD as a result of the convergence of the shock from the He shell detonation and in the few other cases the failure is likely a result of the finite numerical resolution. The high maximum densities and temperatures that are needed for detonation ignition and the fact that the shock is not diluted too much when propagating inwards against the pressure gradient are made possible through a geometrical shock amplification effect that appears in spherical symmetry: As the surface of the shock decreases its strength has to increase. The high compression that can be achieved theoretically in this way, especially if full convergence is reached and the shock is reflected (cf.\\ Sect.~\\ref{sec:res}), could even allow successful double detonations with considerably smaller He masses. The shock amplification that is reached, however, turns out to be resolution dependent on the numerical grid as it is coupled to the smallest resolvable shock surface. Taking this into account it is very likely that all our models would explode, if they were simulated with sufficiently high resolution. The question of whether an incineration could also happen directly at the core--shell interface, was not addressed here. This could of course prevent the spherical shock convergence mechanism from playing a role, at least in parts of the parameter space. Therefore, edge-lit detonations are postponed to a separate study. The complete double detonation simulations that were performed for the models z4.24A and s4.10A resulted in \\nni-masses of about $0.40$ to $0.45~\\msol$. Are the studied events thus good candidates for normal SNe~Ia? Most likely not. Starting with a He-shell detonation, all of our models show a significant amount of rapidly expanding \\nni\\ in the outer layers. In the observed spectra of normal and sub-luminous events, however, this has never been observed (cf.\\ \\citet{Branch1982,Branch1984a,Woosley1986,Arnett1997} and also the discussion in \\citet{Livne1995}). The only exception is the super-luminous \\object{SN~1991T}. Thus our solar-mass models most likely will not be able to explain normal or sub-luminous SNe~Ia and, given the robustness of the explosion mechanism of the model, the only conceivable explanation for the lack of observations of corresponding SN~Ia events is that the progenitors are not realized in nature -- or are very rare. For a better agreement with observations a reduction of the He-shell mass would be the most obvious choice. However, then a significantly larger core mass would most likely be required, because otherwise the He would not detonate. In this case the core density would be much higher resulting in a very large Ni mass. This would again not be a candidate for normal SNe~Ia, but it might be a promising candidate for objects like \\object{SN~1991T}. Further work could cover model improvements like a more realistic treatment of nuclear reactions including the calculation of real reaction rates depending on the actual thermodynamic state of the burnt matter. This would make an investigation of the onset and the explosion dynamics of the core detonation possible. Also, more extended parameter studies, especially towards the lower end of possible He masses, would be interesting. For a comparison with observations the calculation of synthetic light curves and spectra would also be desirable." }, "0710/0710.5023_arXiv.txt": { "abstract": "{} {The principal aim of this project is to determine the structural parameters of the rapidly pulsating subdwarf B star PG 0911+456 from asteroseismology. Our work forms part of an ongoing programme to constrain the internal characteristics of hot B subdwarfs with the long-term goal of differentiating between the various formation scenarios proposed for these objects. So far, a detailed asteroseismic interpretation has been carried out for 6 such pulsators, with apparent success. First comparisons with evolutionary theory look promising, however it is clear that more targets are needed for meaningful statistics to be derived.} {The observational pulsation periods of PG 0911+456 were extracted from rapid time-series photometry using standard Fourier analysis techniques. Supplemented by spectroscopic estimates of the star's mean atmospheric parameters, they were used as a basis for the \"forward modelling\" approach in asteroseismology. The latter culminates in the identification of one or more \"optimal\" models that can accurately reproduce the observed period spectrum. This naturally leads to an identification of the oscillations detected in terms of degree $\\ell$ and radial order $k$, and infers the structural parameters of the target.} {The high S/N low- and medium resolution spectroscopy obtained led to a refinement of the atmospheric parameters for PG 0911+456, the derived values being $T_{\\rm eff}$=31,940$\\pm$220 K, $\\log{g}$=5.767$\\pm$0.029, and $\\log{\\rm He/H}$=$-$2.548$\\pm$0.058. From the photometry it was possible to extract 7 independent pulsation periods in the 150$-$200 s range with amplitudes between 0.05 and 0.8 \\% of the star's mean brightness. There was no indication of fine frequency splitting over the $\\sim$68-day time baseline, suggesting a very slow rotation rate. An asteroseismic search of parameter space identified several models that matched the observed properties of PG 0911+456 well, one of which was isolated as the \"optimal\" model on the basis of spectroscopic and mode identification considerations. All the observed pulsations are identified with low-order acoustic modes with degree indices $\\ell$=0,1,2 and 4, and match the computed periods with a dispersion of only $\\sim$ 0.26 \\%, typical of the asteroseismological studies carried out to date for this type of star. The inferred structural parameters of PG 0911+456 are $T_{\\rm eff}$=31,940$\\pm$220 K (from spectroscopy) , $\\log{g}$=5.777$\\pm$0.002, $M_{\\ast}/M_{\\odot}$=0.39$\\pm$0.01, $\\log{M_{env}/M_{\\ast}}$=$-$4.69$\\pm$0.07, $R/R_{\\odot}$=0.133$\\pm$0.001 and $L/L_{\\odot}$=16.4$\\pm$0.8. We also derive the absolute magnitude $M_V$=4.82$\\pm$0.04 and a distance $d$=930.3$\\pm$27.4 pc.} {} ", "introduction": "Subdwarf B (sdB) stars are evolved extreme horizontal branch stars with atmospheric parameters in the 20,000 K $\\lesssim T_{\\rm eff} \\lesssim$ 40,000 K and 5.0 $\\lesssim \\log{g} \\lesssim$ 6.2 range \\citep{heber1986}. They are believed to be composed of helium-burning cores surrounded by thin hydrogen-rich envelopes and are characterised by a narrow mass distribution strongly peaked at $\\sim$ 0.48 $M_{\\odot}$ \\citep{dorman1993}. While it is generally accepted that they evolved from the red giant branch (RGB) and constitute the immediate progenitors of low-mass white dwarfs \\citep{bergeron1994}, the details of their formation are not yet understood. It does however seem clear that sdB progenitors lost a significant fraction of their envelope mass near the tip of the first RGB, leaving them with insufficient fuel to ascend the asymptotic giant branch (AGB) after core helium exhaustion. Plausible formation channels were modelled in detail by, e.g., \\citet{han2002,han2003} and include binary evolution via a common envelope (CE) phase, stable Roche lobe overflow (RLOF), and the merger of two helium white dwarfs. These distinct evolutionary scenarios should leave a clear imprint not only on the binary distribution of sdB stars (CE evolution will produce sdB's in very close binary systems, RLOF gives rise to much longer period binaries, and the white dwarf merger results in single sdB stars), but also on their mass and hydrogen-envelope thickness distribution. Observational surveys focusing on radial velocity variations and the spectroscopic detection of companions have recently established that at least half of the sdB population reside in binaries (e.g. \\citet{allard1994,ulla1998}), a significant fraction of them having short orbital periods from hours to days (\\citet{green1997,maxted2001}). Accurate determinations of the stars' internal parameters on the other hand are harder to come by using traditional techniques; the mass has so far been measured only for the very rare case of an eclipsing binary \\citep{wood1999}, while the envelope thickness eludes direct study. Fortunately, a small fraction ($\\sim$ 5 $\\%$) of sdB stars have been discovered to exhibit rapid, multi-periodic luminosity variations on a time-scale of hundreds of seconds, thus opening up the possibility of using asteroseismology to constrain their internal parameters. Since the near-simultaneous theoretical prediction \\citep{charp1996, charp1997} and observational discovery \\citep{kilkenny1997} of the so-called EC 14026 stars, both the modelling and measurement of their pulsational properties have come a long way (see \\citet{fontaine2006} for a review). Simulating the pulsation spectra of a large grid of sdB models in terms of low-degree, low-order $p$-modes and numerically determining the \"optimal\" model that best fits a series of observed periodicities has so far resulted in the asteroseismological determination of the internal parameters for six EC 14026 pulsators: PG 0014+067 \\citep{brassard2001}, PG 1047+003 \\citep{charp2003}, PG 1219+534 \\citep{charp2005a}, Feige 48 \\citep{charp2005b}, EC 20117-4014 \\citep{randall2006c}, and PG 1325+101\\citep{charp2006}. These first asteroseismic results show promising trends as far as matching the expected mass and hydrogen shell thickness distribution is concerned, however more targets are needed to start assessing the importance and accuracy of the proposed formation channels. Here we present an asteroseismological analysis of the subdwarf B star PG 0911+456 based on photometry obtained with the new Mont4kccd at Mt. Bigelow, Arizona. In the next sections we describe the observations and frequency analysis, followed by the asteroseismic exploitation of the target and a discussion of the internal parameters inferred. ", "conclusions": "We obtained 57 hours of broad-band time-series photometry as well as high S/N low- and medium resolution time-averaged spectroscopy for the EC 14026 pulsator PG 0911+456. Our observations led to refined estimates of the star's atmospheric parameters and the detection of 7 independent harmonic oscillations, 4 more than were known previously. There was no sign of frequency splitting over the 68-day period during which the photometry was obtained, indicating a slow rotation rate. Fixing the effective temperature to the spectroscopic value and conducting an asteroseismic search in 3-dimensional $\\log{g}-M_{\\ast}-\\log{q(\\rm H)}$ parameter space enabled the identification of several families of models that could reproduce the observed periods to within less than 1 \\%. While some of these were rejected from the outset due to obvious inconsistencies with the spectroscopic estimate of $\\log{g}$ or implausible associated mode identifications, we retained two promising solutions for closer inspection. Unlike in some previous asteroseismological studies, it was not immediately obvious which model was to be preferred on the basis of the structural parameters alone; instead we used the inferred mode identification, in particular the degree index $\\ell$, to discriminate between the two. The main difference between the solutions was that one identified a relatively high amplitude peak with an $\\ell$=3 mode, while the other required only modes with $\\ell$=0,1,2 and 4. Since detailed computations reveal $\\ell$=3 modes to have extremely small disk-integrated amplitudes that would most likely not be detectable, we favoured the latter and adopted it as the optimal model. The inferred structural parameters for PG 0911+456 include the total stellar mass and the thickness of the hydrogen-rich shell, two quantities that can normally not be derived using other means but are invaluable for a detailed understanding of subdwarf B stars' evolutionary history. The total mass determined is smaller than that found for any EC 14026 star to date and places our target at the low-mass end of the predicted distribution. Similarly, the hydrogen envelope is measured to be thinner than that of most sdB's studied so far. If PG 0911+456 is confirmed to be a single star, as is suspected from its slow rotation, negligible radial velocity variation, and absence of a companion's spectroscopic or near-IR photometric signature, it may be the product of a WD merger according to the evolutionary channels proposed by \\citet{han2002,han2003}. In this case, we would indeed expect the hydrogen envelope mass to be smaller than for an sdB having undergone a CE or RLOF phase. The low total mass derived would tend to support a non-canonical evolutionary history, even if it lies slightly below the mass distribution predicted from a WD merger. An obvious follow-up study for PG 0911+456 is to verify the asteroseismological solution found on the basis of additional observations. These could aim for a higher S/N level, thus enabling the detection of further pulsations and strengthening the constraints on the asteroseismic model. One of the main challenges we faced in the search of parameter space was the relatively large number of models that could account for the oscillations observed quite accurately; it would be very instructive to see whether any newly found periods can also be fit by the optimal model isolated. Observations containing wavelength-dependent information from which modes may be partially identified are another option. Following recent theoretical investigations into the amplitude-wavelength dependence of a mode on its degree $\\ell$ (\\citet{ramachandran2004}, \\citet{randall2005}), there has been a surge in observational efforts to obtain multi-colour photometry of EC 14026 stars, most notably using the 3-channel CCD ULTRACAM (e.g. \\citet{jeffery2005}) and the Mont4kccd predecessor, LaPoune I (e.g. \\citet{fontaine2006}). While discriminating between low-degree modes with $\\ell$=0,1,2 has proved extremely challenging, $\\ell$=4 modes exhibit a more clearly distinguishable amplitude-wavelength behaviour (see e.g. Figure 26 of \\citet{randall2005}). Given the necessary data, it should therefore be possible to confirm the identification of the two $\\ell$=4 modes inferred for PG 0911+456 and thus verify the structural parameters computed. The work presented here constitutes the 7th detailed asteroseismological analysis of a rapidly pulsating subdwarf B star. While we estimate that the structural parameters of around 20 targets are required to start detailed comparisons with evolutionary theory, first tentative efforts in this direction look promising. Nevertheless, it is clear that there is still ample room for improvement on the modelling front. Firstly, the fact that the dispersion between the observed and theoretical periods of the optimal model is generally an order of magnitude higher than the measurement accuracy indicates remaining shortcomings in the models. We are currently working on full evolutionary (rather than envelope) \"third-generation\" models to address this problem. The reliability of the \"optimal\" models identified during the search of parameter space is another issue. Although we do apply cross-checks such as compatibility with non-adiabatic theory and spectroscopic values of the atmospheric parameters, the latter (especially $T_{\\rm eff}$) must often be used to discriminate between, or constrain, regions of minimum $S^2$ and can no longer be employed as independent estimates. Moreover, the period ranges of unstable oscillations computed from our non-adiabatic pulsation code are sensitive mostly to $T_{\\rm eff}$ and $\\log{g}$ and are largely independent of the model mass and envelope thickness. It is therefore vital that additional checks are carried out with regard to the robustness of the \"forward\" approach if the structural parameters inferred from asteroseismology are to be compared with evolutionary predictions in a quantitative manner. The most obvious way of doing this is by detecting more frequencies from higher S/N observations or constraining the identification of the degree $\\ell$ of individual modes from multi-wavelength time-series data. Such efforts are ongoing, and will likely prove invaluable for the future of sdB star asteroseismology." }, "0710/0710.2435_arXiv.txt": { "abstract": " ", "introduction": "Magnetic fields are observed to be associated with most structures in the universe. Observations indicate magnetic fields on stellar upto supergalactic scales. The field strengths vary from a few $\\mu$G on galactic scale, upto $10^3$ G for solar type stars and upto $10^{13}$ G for neutron stars. Furthermore, the magnetic field structure depends on the object it is associated with. Thus, e.g., magnetic fields observed in elliptical galaxies show a different structure from those associated with spiral galaxies \\cite{mag}. Magnetic fields in stars can be explained by the formation of protostars out of condensed interstellar matter which was pervaded by a pre-existing large scale magnetic field (see, e.g., \\cite{rees}). An open problem remains to explain the origin of such large scale magnetic fields. There are different types of proposals. Ranging from processes on small scales, such as vortical perturbations and phase transitions to models taking advantage of the possibility of amplifying perturbations in the electromagnetic field during inflation in the early universe (see, e.g., \\cite{rev-mag}). Inflation provides a mechanism to amplify perturbations in some field to appreciable size. In order for this mechanism to lead to primordial magnetic seed fields of cosmologically interesting strength, the corresponding lagrangian should not be conformally invariant. The Maxwell lagrangian describing linear electrodynamics is conformally invariant. There have been already a multitude of proposals to break the conformal invariance of the Maxwell theory \\cite{tw}, e.g. by coupling to a scalar field \\cite{mag-sc}, breaking Lorentz invariance \\cite{mag-lor}, adding extra dimensions \\cite{mag-ex} or a coupling to curvature terms \\cite{mag-curv}. Here nonlinear electrodynamics is considered. It has its origins in the search for a classical singularity-free theory of the electron by Born and Infeld \\cite{bi}. Later on it was realized that virtual electron pair creation induces a self-coupling of the electromagnetic field. For slowly varying, but arbitrarily strong electromagnetic fields the self-interaction energy was computed by Heisenberg and Euler (cf. \\cite{he}-\\cite{bb}). The propagation of a photon in an external electromagnetic field can be described effectively by the Heisenberg-Euler langrangian. Moreover, the transition amplitude for photon splitting in quantum electrodynamics is nonvanishing in this case. In principle, this might lead to observational effects, e.g., on the electromagnetic radiation coming from neutron stars which are known to have strong magnetic fields. \\cite{bb,p-sp}. In particular, certain features in the spectra of pulsars can be explained by photon splitting \\cite{puls}. Finally, Born-Infeld type actions also appear as a low energy effective action of open strings \\cite{bi-strings,gh}. As was shown in \\cite{dbi} the low energy dynamics of D-branes is described by the Dirac-Born-Infeld action. The model of the cosmological background that will be considered consists of a stage of de Sitter inflation followed by reheating and a standard radiation dominated stage. Quantum fluctuations in the electromagnetic field are excited within the horizon during inflation. Once outside the horizon they become classical perturbations. As mentioned above, in general, the conformal invariance of the four dimensional Maxwell field has to be broken in order to amplify the perturbations in the electromagnetic field significantly. Thus, here electrodynamics is considered to be nonlinear during the de Sitter stage. This could be motivated by the presence of possible quantum corrections to quantum electrodynamics at high energies. However, once inflation ends electrodynamics is described by standard Maxwell electrodynamics. Thus the subsequent evolution described by the standard model of cosmology is unchanged. ", "conclusions": "Observations of magnetic fields on large scales provide an intriguing problem. A possible class of mechanisms to create such fields is provided by inflationary models. Fluctuations in the electromagnetic field are amplified during inflation and provide a seed magnetic field at the time of structure formation which might be further amplified by a dynamo process. In general a sufficiently strong initial field strength can only be achieved if the conformal invariance of electrodynamics is broken. This has been realized, for example, in models where the Maxwell lagrangian has been coupled to a scalar field, to curvature terms, etc. or by breaking Lorentz invariance or adding extra dimensions. Here nonlinear electrodynamics has been considered. It has been assumed that whereas during the early universe electrodynamics is nonlinear it becomes linear at the end of inflation. In particular the evolution of the magnetic energy density has been studied in a model of nonlinear electrodynamics which is described by a lagrangian of the form $L\\sim -\\left[\\left(F_{\\mu\\nu}F^{\\mu\\nu}\\right)^2/\\Lambda^8\\right]^{\\frac{\\delta-1}{2}} F_{\\mu\\nu}F^{\\mu\\nu}$, where $\\Lambda$ and $\\delta$ are parameters. Originally the nonabelian version of this model had been proposed to describe low energy QCD \\cite{pt}. Here this model has been chosen as it is a strongly nonlinear theory of electrodynamics which allows to study in a semi-analytical way the possible creation and amplification of a primordial magnetic field during de Sitter inflation. This is so since on the one hand the lagrangian only depends on one of the electromagnetic invariants, namely $X=\\frac{1}{4}F_{\\mu\\nu}F^{\\mu\\nu}$, which leads to a significant simplification of the equations. On the other hand the power-law structure of the lagrangian make the equations simpler. Approximate solutions have been found in three regimes of approximation which describe the evolution of the ratio of the energy densities of the electric and magnetic fields during inflation. It is assumed that initially the energy density of the electric and magnetic field are of the same order. Furthermore, these initial fields are due to quantum fluctuations in the electromagnetic field during inflation. Whereas in the radiation dominated era, the energy density in the magnetic field decreases as $a^{-4}$, the electric field strength rapidly decays in the highly conducting plasma (see, e.g., \\cite{tw,d93}). Solutions in closed form have been found and the resulting primordial magnetic field estimated. It has been shown that depending on the regime of approximation and the value of the Pagels-Tomboulis parameter $\\delta$ primordial magnetic fields can be generated that are strong enough to seed a galactic dynamo. Thus we have provided an example of a theory of nonlinear electrodynamics where the nonlinearities act in a way as to amplify sufficiently an initial magnetic field. The energy-momentum tensor of the electromagnetic field can be cast in the form of an imperfect fluid. This has been found explicitly for the particular model of nonlinear electrodynamics under consideration here. Moreover, this allows to find the expression for the energy density $\\rho$ of the fluid. Requiring that $\\rho$ should be positive provides the bound $\\delta\\geq\\frac{1}{2}$. In \\cite{gfc} the possible creation and amplification of magnetic fields was studied in an inflationary model coupled to a pseudo Goldstone boson (see also \\cite{tw}). In this case the lagrangian has the form $L\\sim\\frac{1}{2}\\partial_{\\mu}\\theta\\partial^{\\mu}\\theta -X+g_a\\theta Y$, where $\\theta$ is the axion field. This provides an example of a more general lagrangian having also an explicit dependence on $Y=\\frac{1}{4}F_{\\mu\\nu}\\;^{*}F^{\\mu\\nu}$. However, as it turns out the resulting primordial magnetic field is not strong enough in order to seed, for example, a galactic dynamo. In \\cite{as} the model of \\cite{gfc} was generalized to N axions. In this case it was found that at least the weaker bound of $r>10^{-57}$ can be satisfied. Here, in this work the creation of primordial magnetic fields in a particular model of nonlinear electrodynamics has been studied. It might be interesting to generalize this to lagrangians depending on both electromagnetic invariants $X$ and $Y$." }, "0710/0710.1600_arXiv.txt": { "abstract": "Prior to the incineration of a white dwarf (WD) that makes a Type Ia supernova (SN Ia), the star ``simmers'' for $\\sim 1000$ years in a convecting, carbon burning region. We have found that weak interactions during this time increase the neutron excess by an amount that depends on the total quantity of carbon burned prior to the explosion. This contribution is in addition to the metallicity ($Z$) dependent neutronization through the $^{22}$Ne abundance (as studied by Timmes, Brown, \\& Truran). The main consequence is that we expect a ``floor'' to the level of neutronization that dominates over the metallicity contribution when $Z/Z_\\odot\\lesssim2/3$, and it can be important for even larger metallicities if substantial energy is lost to neutrinos via the convective Urca process. This would mask any correlations between SN Ia properties and galactic environments at low metallicities. In addition, we show that recent observations of the dependences of SNe Ia on galactic environments make it clear that metallicity alone cannot provide for the full observed diversity of events. ", "introduction": "\\label{sec:introduction} The use of Type Ia supernovae (SNe Ia) as cosmological distance indicators has intensified the need to understand white dwarf (WD) explosions. Of particular importance is the origin of the Phillips relation \\citep{phi99}, an essential luminosity calibrator. Recent models demonstrate that it can be explained by large variations in the abundance of stable iron group elements \\citep{woo07} with the dominant cause for diversity likely residing in the explosion mechanism \\citep{maz07}. One additional variable is the metallicity of the WD core, which yields excess neutrons relative to protons due to the isotope $^{22}$Ne. This is usually expressed as \\be Y_e = \\sum_i \\frac{Z_i}{A_i}X_i, \\ee where $A_i$ and $Z_i$ are the nucleon number and charge of species $i$ with mass fraction $X_i$. The neutronization is critical for setting the production of the neutron-rich isotopes \\citep{thi86}. If no weak interactions occur during the explosion, the mass fraction of $^{56}$Ni produced is simply $X(^{56}{\\rm Ni}) = 58Y_e-28$, assuming $^{56}$Ni and $^{58}$Ni are the only burning products \\citep{tbt03}. The neutronization also affects the explosive burning, including the laminar flame speed \\citep{cha07}. However, the metallicity range of progenitors is not large enough to account for the full SNe Ia diversity (see \\S 4), making it critical to explore all factors that determine $Y_e$. A potential neutronization site is the convective carbon burning core that is active for $\\sim1000\\ {\\rm years}$ prior to the explosion. The hydrostatic evolution associated with this simmering phase terminates when the core temperature is sufficiently high that burning is dynamical \\citep{nom84,ww86,woo04,ww04,kwg06}, and a flame commences \\citep{tw92}. We show here that protons from the $^{12}$C($^{12}$C,$p$)$^{23}$Na reaction during simmering capture on $^{12}$C, and that subsequent electron captures on $^{13}$N and $^{23}$Na decrease $Y_e$. This process continues until either the explosion occurs or sufficient heavy elements have been synthesized that they capture the protons instead. In \\S \\ref{sec:rates}, we present simmering WD core models and summarize the reaction chains that alter $Y_e$. We find that one proton is converted to a neutron for each six $^{12}$C nuclei consumed for burning at $\\rho<1.7\\times 10^9\\ {\\rm g \\ cm^{-3}}$. At densities above this, an additional conversion occurs from an electron capture on $^{23}$Na. Hence, the $Y_e$ in the core depends on the amount of carbon burned during simmering and the density at which it occurs, which we quantify in \\S 3. We find that neutronization during simmering dominates for metallicities $Z/Z_\\odot\\lesssim2/3$. We conclude in \\S \\ref{sec:conclusion} by discussing the observations and noting where future work is needed. ", "conclusions": "\\label{sec:conclusion} We have found that significant neutronization of the WD core occurs throughout the simmering stage of carbon burning until the onset of the explosion. If substantial energy is lost to the convective Urca process \\citep[][and references therein]{les05}, then the neutronization is truncated by proton captures onto freshly synthesized heavy elements (resulting in eq. [\\ref{eq:yemax}]). The main consequence is a uniform ``floor'' to the amount of neutronization that dominates over the metallicity dependent contribution for all progenitors with $Z/Z_\\odot\\lesssim2/3$. Given the likely significance this has for SNe Ia, more work needs to be done. In particular, at high ignition densities, heavy element electron captures and a full reaction network are needed to follow the resulting diverse collection of elements (see the discussion in \\S 2.2). The convective Urca process % is another complication we have not addressed. In principle, if more energy is lost to neutrinos then more burning (and thus more neutronization) is required to make the burning dynamical. Assessing this will necessitate coupling a full nuclear network \\citep{cha07b} to convective calculations. Such models would accurately determine $\\eta$ rather than simply setting it to 1 or 2. In closing, we highlight some important features exhibited by recent observations of SNe Ia. It is clear that the amount of $^{56}$Ni produced in SNe Ia has a dynamic range ($0.1-1M_\\odot$) larger than can be explained by metallicity or simmering neutronization. However, since an intriguing trend is the prevalence of $^{56}$Ni rich events in star-forming regions it is interesting to quantitatively explore how large the observed discrepancy is. Using the SNLS sample of Sullivan et al. (2006), Howell et al. (2007) found that the average stretch is $s=0.95$ in passive galaxies (e.g. E/S0's) and $s=1.05$ in star-forming galaxies. Using Jha et al's (2006) $\\Delta M_{15}(B)-s$ relation and Mazzali et al.'s (2007) relation between $\\Delta M_{15}(B)$ and $^{56}$Ni mass we get $0.58M_\\odot$ ($s=0.95$) and $0.72M_\\odot$ ($s=1.05$). Hence, amongst the large diversity, there is a tendency for SNe in star-forming galaxies to produce $\\approx 0.13M_\\odot$ more $^{56}$Ni than those in large ellipticals. Since the SN Ia rate scales with mass in ellipticals and star formation rate in spirals (Mannucci et al. 2005; Scannapieco \\& Bildsten 2005; Sullivan et al. 2006), SNe from passive galaxies in the SNLS survey are from more massive galaxies than the SNe in star-forming galaxies (Sullivan et al. 2006). Using the mass-metallicity relation of Tremonti et al. (2004), our integration of the separate samples in Sullivan et al. (2006) yield average $12+\\log({\\rm O/H})=8.87$ in active galaxies and $9.1$ in ellipticals (solar value is $12+\\log({\\rm O/H})=8.69$). Due to the increase of ${\\rm N/O}$ at high metallicities (Liang et al. 2006), the SNe in ellipticals have twice as much $^{22}$Ne content as those in spirals. From the relation of Timmes et al. (2003), this implies $\\approx 5\\%$ less $^{56}$Ni, whereas the observed decrement is $>15\\%$. Explaining the observed decrement would require a contrast of $\\Delta X(^{22}$Ne$)\\approx 0.06$, or nearly 3 times solar. Although we have found that simmering enhances neutronization, the effect is not great enough ($\\Delta Y_{e,\\rm max}$ would give the same change in neutronization as doubling a solar metallicity), and a diverse set of core conditions would still be required. A large enhancement could be present in the core if substantial gravitational separation had occurred \\citep{bh01,db02}, yet convective mixing during simmering will reduce it based on the fraction of the star that is convective. For a convection zone that extends out to $M_\\odot$, the resulting $^{22}$Ne enhancement would be at most $\\approx 30\\%$." }, "0710/0710.1752_arXiv.txt": { "abstract": "The early evolution of dense stellar systems is governed by massive single star and binary evolution. Core collapse of dense massive star clusters can lead to the formation of very massive objects through stellar collisions ($M\\geq$ 1000\\,\\msun). Stellar wind mass loss determines the evolution and final fate of these objects, and decides upon whether they form black holes (with stellar or intermediate mass) or explode as pair instability supernovae, leaving no remnant. We present a computationaly inexpensive evolutionary scheme for very massive stars that can readily be implemented in an N-body code. Using our new N-body code 'Youngbody' which includes a detailed treatment of massive stars as well as this new scheme for very massive stars, we discuss the formation of intermediate mass and stellar mass black holes in young starburst regions. A more detailed account of these results can be found in \\cite{Belkusetal}. ", "introduction": " ", "conclusions": "" }, "0710/0710.4946_arXiv.txt": { "abstract": "Thermal instability of partially ionized plasma is investigated by means of a linear perturbation analysis. According to the previous studies under the one fluid approach, the thermal instability is suppressed due to the magnetic pressure. However, the previous studies did not precisely consider the effect of the ion-neutral friction, since they did not treat the flow as two fluid which is composed of ions and neutrals. Then, we revisit the effect of the ion-neutral friction of the two fluid to the growth of the thermal instability. According to our study, the characteristic features of the instability are the following four points: (1) The instability which is characterized by the mean molecular weight of neutrals is suppressed via the ion-neutral friction only when the magnetic field and the friction are sufficiently strong. The suppression owing to the friction occurs even along the field line. If the magnetic field and the friction are not so strong, the instability is not stabilized. (2) The effect of the friction and the magnetic field is mainly reduction of the growth rate of the thermal instability of weakly ionized plasma. (3) The effect of friction does not affect the critical wavelength $\\lambda _{\\rm F}$ for the thermal instability. This yields that $\\lambda _{\\rm F}$ of the weakly ionized plasma is not enlarged even when the magnetic field exists. We insist that the thermal instability of the weakly ionized plasma in the magnetic field can grow up even at the small length scale where the instability under the assumption of the one fluid plasma can not grow owing to the stabilization by the magnetic field. (4) The wavelength of the maximum growth rate of the instability shifts shortward according to the decrement of the growth rate. This is because the friction is effective at rather larger scale. Therefore, smaller structures are expected to appear than those without the ion-neutral friction. Our results indicate the friction with the magnetic field affects the morphology and evolution of the interstellar matter. In summary, the ion-neutral friction is important for the evaluation of the thermal instability in weakly ionized plasma along and perpendicular to the magnetic field. ", "introduction": "The recent progress of the interstellar medium (ISM) observations has established that the small and tiny scale structures of ISM is very ubiquitous. Historically, the initial observational target is the neutral phase of ISM. Indeed, the tiny scale structures are discovered by 21 cm absorption line observations against quasars with VLBI techniques \\citep{Di76}. The results are confirmed by \\citet{Dia89}. To look for the small scale structures in the cold neutral phase of the ISM furthermore, the observations of 21 cm absorption lines have been also performed against pulsars \\citep{Fr94}. \\citet{Me96} observe optical spectral lines to find small structures against close binary stars. The images of the small-scale H \\textsc{i} have been taken with the MERLIN array \\citep{Da96}. These results are established by higher angular resolution observations with VLBA and VLA toward quasars \\citep{Fa98,FG01}, which show small clumps on the order of a few AU in neutral Galactic H \\textsc{i} clouds. Cold H \\textsc{i} clouds have significant structure in subparsec scales \\citep{Gi00,Br05}. Such small and tiny scale structures has been detected in the local interstellar medium as H \\textsc{i} absorption lines, although their column density is very small \\citep{BK05}. In addition to the small and tiny scale structures of H {\\rm \\textsc{i}} cloud, the picture of the planetary nebula NGC 7293 shows fine small structures and knots \\citep{Ro02,OD04}. Even in the starforming regions, there are variety small scale structures. For example, \\citet{La95} observe some clumps with size from 0.007 to 0.021 pc in the Taurus Molecular Cloud 1 (TMC1), in particular, core D. The mass of these fragments is estimated to be $\\lesssim 0.01 - 0.15$ $M_{\\odot}$. SPITZER has begun producing higher spatial resolution mid-infrared maps \\citep{Ch04}, and revealed the fine structure of the starforming region. We think that it is possible for proto-brown dwarfs of $< 0.08$ $M_{\\odot}$ to exist there. A high resolution observation in future is expected to reveal hidden, small structures, as well as substellar objects with very small mass. In any ways, the tiny-scale structure seems to be ubiquitous, not associated with large extinction \\citep{He97}. To study the origin of the small-scale structures of ISM is important to understand the evolution of and structure formation in ISM. Especially, in the starforming regions, the small, low-mass structures can relate to the low-mass cutoff of the initial mass function, the coagulation unit to form the massive stars, and so on. Then, we should investigate the physical origin of these small and tiny structures in the partially ionized medium, since the cold H {\\rm \\textsc{i}} and molecular clouds are weakly ionized. By the way, it is not easy for the small and tiny scale structure to form as a result of gravitational instability since their size is much smaller than the Jeans length. According to \\citet{La95}, indeed, the small-scale structures appear to be gravitationally unbound. This suggests that some fragmentation mechanisms but pure Jeans gravitational instability may be important in the clouds. We expect the thermal instability as this mechanism, and revisit it in this paper. The basics of the thermal instability has been summarized in \\citet{F65}. When the following condition is satisfied, the system is thermally unstable: If the cooling becomes efficient as the temperature decreases, the thermal energy gets lost more and more. If the cooling becomes efficient during the fluid contraction, the system shrinks on and on. Importantly, the critical length scale of the thermal instability is smaller than that of the dynamical instability like the Jeans instability. That is, even if a system is stable against the gravitational instability, the system can become thermally unstable. Then, the thermal instability can be the physical origin of the smaller-scale formation than the dynamical instability. Indeed, \\citet{KI02} propose that the clumpiness in clouds emerges naturally from their formation through the thermal instability. Their two-dimensional calculations follow the fragmentation into small cloudlets that result from the thermal instability in a shock-compressed layer. \\citet{BL00} investigate the cooling and fragmentation of optically thin gas with a power-law cooling function. According to them, small-scale perturbations have the potential to reach higher amplitude than large-scale fluctuations. Thermal instability can be important for the structure formation. We sketch ISM structure formation such as molecular clouds as follows. The hot ionized ISM cools to be cold and weakly ionized clouds. Ionization degree of ISM decreases as it evolves, and then the ISM fluid is often partially ionized. In partially ionized fluid, an ion component and neutral component interact each other, exchanging the momentum (i.e., ion-neutral friction). Especially, the partially ionized plasma in a magnetic field could not be treated as one fluid. This is because although the ion component in the partially ionized plasma is directly influenced by a magnetic field, the neutral component does not directly feel a magnetic field. In weakly ionized fluid, the neutral component is affected by a magnetic field via the ion-neutral friction. For example, ambipolar diffusion takes an important role in dynamical evolution owing to the gravitational instability of ISM into protostar \\citep{MeSp56, Na76}. If the fluid frozen in a magnetic field contracts without ambipolar diffusion, the magnetic field becomes too large and suppresses the growth by the magnetic pressure and tension. In addition, \\citet{Na79} points out the possibility that ISM with sufficient magnetic flux quasistatically evolves into protostar. He also notices that the time scale of plasma drift depends on the ionization degree. The relation among the ion-neutral friction, the magnetic field, and the amount of the ion component is important to understand the detailed process. We here emphasize that the MHD approximation is not always applicable for our purpose in this paper. The thermal instability is effective in small structure formation, during which a neutral component is not always enough frozen in an ion component dynamically to be treated as one fluid. There is a possibility that the ion component can drift away owing to ambipolar diffusion, and that the system can contract owing to thermal instability. Then, in the present paper, to study how the ion-neutral drag and magnetic field influence the growth of the thermal instability, we treat the plasma as the two fluid of ions and neutrals. We assume that the ionization degree is very small, because we are interested in the final stage of the formation of a molecular cloud, in which the ionization degree decreases very much. We focus on the condensation mode of thermal instability, since it is important in structure formation, rather than the oscillation (overstable) mode. In \\S2, the problem is formulated. The property of the dispersion relation of the instability is presented in \\S3. Several discussions relating to the instability are found in \\S4. Applicability of our results to cold H {\\rm \\textsc{i}} medium is briefly examined in \\S5, and then we summarize the paper in \\S6. ", "conclusions": "\\label{S_Discuss} We study the basic property of the thermal instability of the weakly ionized plasma in the previous sections. In this section, basing on our understanding the thermal instability, we shall discuss the structure formation. \\subsection{Critical Wavelength and Thermal Conduction} \\label{subsec:critical_wavelength_conduction} First in this section, the critical wavelength of the weakly ionized plasma in the magnetic field is discussed. As mentioned in section \\ref{S_Results}, the critical wavelength of the mode relating $\\mu_{\\rm n}$ is not changed by the magnetic field with the friction as long as the thermal conduction is not changed by the magnetic field. In fact, the thermal conductivity is varied owing to the magnetic field. In this meaning, we may oversimplify the problem in this paper. However, the thermal conductivity will decrease as the magnetic field strength increases, and then the critical length become smaller, according to equation (\\ref{eq:appendix:criticallength:neutral}). Thermal conduction erases the temperature perturbation and thermally stabilizes the system, especially at smaller scale. The larger thermal conduction makes the critical wavelength longer, owing to the criterion \\ref{crite_cond} or \\ref{crite_ion}, according to \\citet{F65}. On the other hand, the magnetic field lessens the thermal conduction of the ion component, because the ion component winds around a magnetic field. Therefore, we can insist that the critical length for the thermal instability of the weakly ionized plasma can not be enlarged by the friction and the magnetic field at least, while the effect of the magnetic field can influence not only the ion but the neutral component via the ion-neutral friction. This statement has universal validity for the thermal instability of the weakly ionized plasma. The effect of the friction and the magnetic field is mainly reduction of the growth rate of the thermal instability. On the other hand, the magnetic field directly enlarges the critical wavelength of the mode characterized by $\\mu_{\\rm i}$ as expected in the one fluid plasma model. \\subsection{Spatial Distribution of Magnetized Interstellar Medium} In the previous section, we learn that the suppression of the thermal instability is characterized by the ion-neutral friction, strength of the magnetic field, and the direction of the field. Interestingly, the suppression of the instability is different owing to the magnetic field direction. This suggests the anisotropic nature of the suppression of the instability is imprinted in the morphology of the structure. In this standing point, we shall discuss how the structure formation proceeds in the magnetized ISM. \\subsubsection{Growth Rate Parallel to the Magnetic Field} In an usual one fluid approach, the motion of the plasma parallel to the magnetic field is regarded not to be affected. However, Figure \\ref{B0C003} suggests that the growth rate of the thermal instability at larger scale decreases via the ion-neutral friction. This suppression of the instability yields the delay of the evolution in comparison to the evolution under the MHD approximation without the ion-neutral friction. \\subsubsection{Distribution of Partially Ionized Plasma Perpendicular to the Magnetic Field Line} \\label{sss_spread_of_part} Here, we suggest that the spatial distribution of the partially ionized plasma is elongated, whose semi-major axis is vertical to the magnetic field line. Figures \\ref{B0C003} and \\ref{B06C003}, both of which friction strengths are the same, show the feasibility. The thermal instability in Figure \\ref{B0C003} corresponds to the mode along the magnetic field line since $\\mbox{\\boldmath$B$} = 0$, while that in Figure \\ref{B06C003} is vertical to the magnetic field line. The growth of thermal instability vertical to the line is suppressed by both of the magnetic field and the friction, while the growth along the line is suppressed only by the friction. Thus, the growth rate along the magnetic field line is larger than that vertical to the line, and then the resultant morphology of the structure via the thermal instability is elongated. \\subsubsection{Difference of Spatial Distribution of Ion and Neutral Components} \\label{ssS_Difference_in_Spread} We insist the growth rate depending on the direction of the field results in the morphology of interstellar structure. Furthermore, we learn in \\S3 the mode characterized by $\\mu_{\\rm i}$ are stabilized by the friction and field, while that by $\\mu_{\\rm n}$ can grow as found in Figure \\ref{B06C003}. Figure \\ref{B600C5} shows that although the mode characterized by $\\mu_{\\rm i}$ is completely stabilized, the mode characterized by $\\mu_{\\rm n}$ is still unstable. Consequently, we can insist that the neutral component condenses owing to the thermal instability, while the ion component still keeps spreading. It is noticed that the ion component may condense since it is dragged by the neutral component contracting owing to the thermal instability. However, once the distribution of the ion and the neutral components separates and the motion of the two becomes independent there, the ion component in the area is thermally stable owing to the magnetic field. This means the magnetic pressure lets the ion component be spread in the area even if the neutral component condenses. This thin ion component left behind by the contracting neutral component will be observed as a halo around H {\\rm \\textsc{i}} clouds, and it emits, especially some forbidden lines. This halo possibly has wide distribution vertical to the magnetic field, as discussed in subsubsection \\ref{sss_spread_of_part}. This is because the ion component can condense, parallel to the magnetic field like the neutral. The magnetic pressure can not prevent the motion of the ion component efficiently as found in Figure \\ref{B0C003}. Thus, the morphology of the ion halo is also elongated according to the orientation of the field line. \\subsubsection{Origin of the Small Clumpinesses of Partially Ionized Plasma} \\label{subsubsec:small_clumpinesses_of_partial} It is suggested that there are small clumpy structures of ISM. Here, we examine whether the clumpiness originates via the thermal instability. According to Figures \\ref{B600C03} and \\ref{B600C5}, the growth rate of the thermal instability of partially ionized plasma depends on the strength of the friction. The scale length at the maximum growth rate of the instability becomes smaller, owing to the suppression via the friction which is effective at larger scale. Thus, we can expect the small clumpiness appears selectively via the thermal instability since the ion-neutral friction suppresses the large scale condensations. If we are interested in the small clumpiness of ISM, we should mention about the thermal conduction in relation to the magnetic field. According to subsection \\ref{subsec:critical_wavelength_conduction}, the small clumpiness may the results of the decrement of the conduction. It is noticed that the conduction along the field line is not so affected, and then there is still ambiguity for the reduced conduction coefficient to perform the key role for the origin of the small clumpiness. It is also noticed that the thermal instability is a slow process relative to the dynamical processes. If molecular clouds form from the H {\\rm \\textsc{i}} medium via the gravitational instability, the small clumpiness inside the molecular clouds appears after the formation of the clouds. This trend is enhanced if the effects of the magnetic field and friction are concerned. This is because the friction and the field suppress the large scale growth, while the short wavelength modes are still unstable. In a enough long span for the growth, the tiny structures can appear in ISM and affect the evolution of its structure especially as the non-linear effects. Thus, if we want to know how faint and small structures come to be observable, the growth time-scale of the thermal instability should be precisely examined. At this time, we never forget the effect of the magnetic field and the ion-neutral friction. \\subsection{Ineffectiveness of the Magnetic Field} \\label{S_Discuss_Satu} The modes in Figures \\ref{B600C03} and \\ref{B6000C03} have similar growth rate, while the magnetic fields are significantly different in each other. It implies that the suppression of the growth by increasing the magnetic field with the fixed friction has limits. By the way, a dispersion relation of only ionized fluid in a magnetic field, which is derived by equalizing the second bracket on the left-hand side in equation (\\ref{dis-rela}) to be zero, is reduced to \\begin{equation} n^3 + n^2 v_{\\rm si} k_{\\rm iT} = 0 , \\end{equation} when $k$ approaches $0$. This is solved to be \\begin{equation} n=0 ~ {\\rm or}~ n = -v_{\\rm si} k_{\\rm iT} . \\end{equation} Thus, we find that this solution is independent of a magnetic field. Even if the magnetic field becomes strong, the value of the growth rate at $k \\sim 0$ is scarcely changed, where the friction is more effective than at small wavelength. Therefore, once the effect of the friction is fixed, increasing the magnetic field strength beyond a threshold level apparently becomes ineffective for the suppression of the growth. Not only information of the magnetic field but the friction is necessary to investigate the evolution of the thermal instability of the weakly ionized plasma. \\subsection{Comment on One Fluid Approximation} The formulation of the fully one component plasma can not be always applicable to study the thermal instability of weekly ionized plasma. This is because there is the growing mode even when there exist the magnetic field and friction, whose growth rate is comparable to that of the neutral component when there are no magnetic field and friction. Indeed, this trend is found from the comparison among Figures \\ref{B06C003}, \\ref{B600C03} and \\ref{B6000C03}. In addition, regarding weakly ionized plasma including the friction effect, the friction does not change the critical wavelength $\\lambda_{\\rm F}$. The friction tends to efficiently reduce the growth at much larger scale than the $\\lambda_{\\rm F}$. This feature yields that the critical wavelength in the weakly ionized plasma is not changed by the magnetic field via the friction, as mentioned in section \\ref{S_Results} and subsection \\ref{subsec:critical_wavelength_conduction}. Even if the magnetic field becomes strong, only the growth rate is reduced, keeping the critical wavelength constant. This is also the difference from the simple analysis of the one component magnetized plasma. The previous study of the one component plasma shows the stabilization at the smaller scales, while the current study insist $\\lambda_{\\rm F}$ is unchanged even if the effect of the field is concerned. \\label{S_Summary} The thermal instability of the weakly ionized fluid is investigated with a linear perturbation analysis. The plasma is assumed to consist of two fluids of the ion and the neutral components. With this approach, the effect of the friction between the ion and the neutral components and the magnetic field are important. Here, the properties of the thermal instability of weakly ionized plasma and the observational implications are summarized. \\begin{enumerate} \\item Modes relating to $\\mu_{\\rm n}$ are not stabilized by the magnetic field. This is because the neutral component feels the field via the ion-neutral friction. \\item Modes relating to $\\mu_{\\rm i}$ are directly stabilized by the magnetic field. This means that when the magnetic field is large, the ion component is hard to follow the motion of the neutrals which condensate owing to the thermal instability. \\item The growth rate of the mode vertical to the magnetic field is reduced by the magnetic field and the ion-neutral friction. \\item The growth rate of the mode along the magnetic field decreases owing to the ion-neutral friction, especially at large scale. The suppression of the instability is more ineffective than that of the mode vertical to the magnetic field. \\item The critical wavelength for the thermal instability is not affected by the friction. The effect owing to the two fluid approximation is the reduction of the growth rate of the thermal instability. The magnetic field makes the critical wavelength of modes relating $\\mu_{\\rm i}$ larger. \\item To study the thermal instability of the partially ionized plasma, one fluid approximation is not always useful. \\item The ion-neutral friction and the magnetic field affect the distribution or morphology of ISM, especially after the long time compared to the free-fall time. The difference between the growth rate along and perpendicular to the magnetic field is important. The partially ionized plasma possibly is elongated perpendicular to the magnetic field. In addition, the neutral component and the ion component of weakly ionized fluid in the magnetic field are possibly separated each other. Then, the neutral component condenses owing to thermal instability, while the ion component left behind by the contracting neutral component still keeps spreading. Therefore, it is possibly observed that the weakly ionized fluid is elongated vertically to the magnetic field and surrounded by a halo which is rare and diffuse ions and emits forbidden lines. \\end{enumerate} To conclude, the treating both of the independent motions of the neutral and the ion components yield the ion-neutral friction. The friction is important for the thermal instability of weakly ionized fluid, especially when studying the structure formation such as molecular cloud at its final stage of formation. Our study indicates that the fully ionized plasma approximation or totally neutral fluid approximation is not always applicable to weakly ionized plasma." }, "0710/0710.1564_arXiv.txt": { "abstract": "{% In a self-absorbed synchrotron source with power-law electrons, rapid inverse Compton cooling sets in when the brightness temperature of the source reaches $T_{\\rm B}\\sim10^{12}\\,$K. However, brightness temperatures inferred from observations of intra-day variable sources (IDV) are well above the \"Compton catastrophe\" limit. This can be understood if the underlying electron distribution cuts off at low energy.} {% We examine the compatibility of the synchrotron and inverse Compton emission of an electron distribution with low-energy cut-off with that of IDV sources, using the observed spectral energy distribution of S5~0716+714 as an example.} {% We compute the synchrotron self-Compton (SSC) spectrum of monoenergetic electrons and compare it to the observed spectral energy distribution (SED) of S5~0716+714. The hard radio spectrum is well-fitted by this model, and the optical data can be accommodated by a power-law extension to the electron spectrum. We therefore examine the scenario of an injection of electrons, which is a double power law in energy, with a hard low-energy component that does not contribute to the synchrotron opacity.} {% We show that the double power-law injection model is in good agreement with the observed SED of S5~0716+714. For intrinsic variability, we find that a Doppler factor of $\\doppler\\geq30$ can explain the observed SED provided that low-frequency ($<32\\,$GHz) emission originates from a larger region than the higher-frequency emission. To fit the entire spectrum, $\\doppler\\ge65$ is needed. We find the constraint imposed by induced Compton scattering at high $T_{\\rm B}$ is insignificant in our model. } {% We confirm that electron distribution with a low-energy cut-off can explain the high brightness temperature in compact radio sources. We show that synchrotron spectrum from such distributions naturally accounts for the observed hard radio continuum with a softer optical component, without the need for an inhomogeneous source. The required low energy electron distribution is compatible with a relativistic Maxwellian.} ", "introduction": "\\label{intro} Observations of many extra-galactic radio sources have found rapid flux variations at radio frequency \\citep[e.g.][]{kedziorachudczeretal01}, some of which fluctuate over a time scale of a day or less. They are referred to as intra-day variable sources (IDV). The variability time scale is often used to constrain the size of the source based on causality arguments. Using this constraint, one can derive a variability brightness temperature \\citep{wagnerwitzel95} \\eqb T_{\\rm var}=4.5\\times10^{10}F_{\\nu}\\left({\\lambda d_{\\rm L}\\over t_{\\rm obs} (1+z)}\\right)^2\\, {\\rm K} \\eqe where the flux density $F_\\nu$, wavelength $\\lambda$, luminosity distant $d_{\\rm L}$, and observed variability time scale $t_{\\rm obs}$ are measured in Jy, cm, Mpc, and days, respectively. The high radio flux frequently measured in IDV sources implies an extremely high brightness temperature, often many orders of magnitude above $10^{12}\\,$K. \\citet{kellermannpaulinytoth69} have shown that, assuming the electron distribution follows a single power law, the luminosity of the inverse Compton scattered photons exceeds that of the synchrotron photons when the brightness temperature of the source reaches $\\sim10^{12}\\,$K. Above this threshold, rapid cooling of the relativistic electrons due to inverse Compton scattering --- the \\lq\\lq Compton catastrophe\\rq\\rq\\ --- forbids a further increase in the brightness temperature \\citep[see e.g.][for a recent review of the brightness temperature problem]{kellermann02}. The limiting value is even lower, $T_{\\rm B}<10^{11}\\,$K, if the magnetic field and particle energy density of the source is driven towards equipartition \\citep{readhead94}. The observed variability in some sources can be interpreted as the result of extrinsic effects, which, at first sight, relaxes the size constraint. For example, the flux variations of PKS~1519$-$273 and PKS~0405$-$385 are convincingly identified as interstellar scintillation. Nevertheless, all realistic models of the scintillation mechanism impose a new constraint on the size and require a brightness temperature of $T_{\\rm B}>10^{13}\\,$K in some cases \\citep{macquartetal00,rickettetal02}, far exceeding the limit imposed by the Compton catastrophe. A prevalent feature associated with IDV sources is a flat or inverted spectrum ($\\alpha\\leq0$, with flux $F_\\nu\\propto\\nu^{-\\alpha}$) at radio-millimeter wavelengths \\citep[e.g.,][]{gearetal94,kedziorachudczeretal01}. Optically thick synchrotron emission from power-law electrons rises as $\\nu^{5/2}$, too fast to account for the observed spectra. Optically thin synchrotron emission in the scope of the conventional interpretation of the synchrotron theory has a flux $F_\\nu\\propto\\nu^{-(s-1)/2}$, where $s$ is the power-law index of the electrons ($\\diff\\nelec/\\diff\\gamma\\propto\\gamma^{-s}$). If $\\alpha=(s-1)/2\\leq0$, the number density of electrons diverges towards high $\\gamma$. Imposing a high-energy cut-off in the electron spectrum avoids the divergence and may account for the commonly observed spectral steepening at optical frequencies, but \\citet{marscher77} showed that electron spectra with $s\\leq1$ would result in a high flux between infrared and optical frequencies that is not supported by observations. The most common interpretation of the flat or inverted spectra is, therefore, a superposition of many synchrotron spectra within an inhomogeneous source \\citep[e.g.][]{debruyn76,marscher77,blandfordkoenigl79}. In \\cite{kirktsang06}, we discussed a synchrotron self-Compton model in which the electron distribution is monoenergetic. The lack of low-energy electrons enables more GHz photons to emerge from the source, allowing a higher brightness temperature to be observed without initiating catastrophic cooling. We found that a temperature of up to $T_{\\rm B}\\sim10^{14}\\,$K at GHz frequencies is possible with only a moderate Doppler boosting factor of $\\sim10$. In \\citet{tsangkirk07}, we discussed the parameters of the monoenergetic model and showed that the assumption of equipartition of energy in the source does not prevent the Compton catastrophe. We also showed that an injection of highly relativistic electrons or strong acceleration in the source cannot produce temperatures much higher than our limit due to copious electron-positron pair production. In this paper, we examine the spectral properties of synchrotron emission from monoenergetic electrons and from an electron distribution that is a double power law in energy, by comparing the model spectra with the observations of S5~0716+714, a BL~Lac object that is one of the brightest known IDV sources, as well as a gamma-ray blazar \\citep{hartmanetal99}. In doing so, we assume that the dominant targets for inverse Compton scattering are produced within the source (SSC model). The emission from gamma-ray blazars can also be interpreted in the context of models in which the target photons are created externally (EC model), for example in the broad line region, the accretion disk, or a molecular torus \\citep{sokolovmarscher05}. However, in many sources there is no observational evidence of a significant external photon source. This is the case for S5~0716+714, where, despite much effort over the past three decades, no emission lines have been detected \\citep[e.g.,][]{bychkova06}. Furthermore, XMM-Newton observations of S5~0716+714 in 2004 analysed by \\citet{ferreroetal06} and \\citet{foschinietal06} show two spectral components in the $0.5-10\\,$keV band, whose variability properties appear to favour the SSC interpretation. The recent extensive simultaneous observations of this object from radio to optical frequencies by \\citet{ostoreroetal06}, together with INTEGRAL pointings at GeV $\\gamma$-ray energies during the same period, provide the best test for our model. In the following, we present the computation of the stationary electron distribution and the resulting synchrotron and inverse Compton spectra. The model spectra computed using the monoenergetic electron approximation, as described in \\citet{tsangkirk07}, are presented first. Although adequate for the radio emission, the monoenergetic model cannot reproduce the entire spectrum of S5~0716+714. We therefore investigate an electron distribution that is a double power law in energy --- a hard low-energy part that softens to a high-energy tail above a characteristic energy. In this way, the inverted optically thin radio emission is retained and complemented by nonthermal synchrotron emission from the high energy tail. In section~\\ref{parameter}, we briefly describe these injection models. The resulting stationary electron distribution is calculated in section~\\ref{stationarysoln} and used for the computation of the synchrotron and inverse Compton spectra. In section~\\ref{sed}, we compare the predictions of these models with the observed spectral energy distribution (SED) of the source to S5~0716+714. Our findings and some limitations of our approach are discussed in section~\\ref{discussion} and our conclusions presented in section~\\ref{conclusion}. ", "conclusions": "\\label{conclusion} Using the specific case of S5~0716+714 as an example, we confirm that it is possible to produce high brightness temperatures at GHz frequencies in compact radio sources without the onset of catastrophic cooling, provided that the radiating particles have a distribution that is sufficiently hard below a characteristic. In addition, we show qualitatively that induced Compton scattering is insignificant in sources with a low-energy electron cut-off despite the high brightness temperature, the underlying reason being the low occupation number of the photons that can couple with the electrons at the cut-off energy. The model where an electron distribution that is a double power law in energy, peaking at $\\gammap$, is injected into the source offers more flexibility at higher frequencies in the synchrotron spectrum (from infrared to optical) at the expense of more free parameters, compared to either monoenergetic or single power-law distributions. These parameters should be constrained by simultaneous observations due to the highly variable nature of IDV sources. In the case of S5~0716+714 where such data is available, the spectral break at about $230\\,$GHz determines the value of $\\gammap$, the optical data at $5\\times10^{14}\\,$Hz gives the lower limit of $\\numax$, as well as constraining the spectral index $s_2$, and the INTEGRAL upper limits give the upper limit of $\\numax$ and also constrain the value of $r_{\\rm p}$, which in turn determines the electron density. The example of S5~0716+714 illustrates several important spectral properties of an electron distribution with a low-energy cut-off, as described in the previous sections. The most noticeable feature is the hard, inverted, optically thin synchrotron spectrum, spanning a wide frequency range, which is a prevalent feature in compact radio sources at radio frequencies \\citep[e.g.,][]{gearetal94,kedziorachudczeretal01}. Other features are the spectral breaks at $\\nup=\\gammap^2\\nu_0$, $\\nucool=\\gcool^2\\nu_0$, and the exponential cut-off at $\\numax=\\gammamax^2\\nu_0$. This model, therefore, allows a simple homogeneous source to reproduce the common features shown by many IDV sources. \\begin{acknowledgement} We thank Luisa Ostorero and Stefan Wagner for helpful discussions and for providing us with easy access to the observational data. We would also like to thank the anonymous referee for constructive comments and suggestions that we feel have led to a significant improvement in this paper. \\end{acknowledgement}" }, "0710/0710.1939_arXiv.txt": { "abstract": "We present a new code aimed at the simulation of diffusive shock acceleration (DSA), and discuss various test cases which demonstrate its ability to study DSA in its full time-dependent and non-linear developments. We present the numerical methods implemented, coupling the hydrodynamical evolution of a parallel shock (in one space dimension) and the kinetic transport of the cosmic-rays (CR) distribution function (in one momentum dimension), as first done by Falle. Following Kang and Jones and collaborators, we show how the \\emph{adaptive mesh refinement} technique (AMR) greatly helps accommodating the extremely demanding numerical resolution requirements of realistic (Bohm-like) CR diffusion coefficients. We also present the \\emph{parallelization} of the code, which allows us to run many successive shocks at the cost of a single shock, and thus to present the first direct numerical simulations of linear and non-linear \\emph{multiple} DSA, a mechanism of interest in various astrophysical environments such as superbubbles, galaxy clusters and early cosmological flows. ", "introduction": "Diffusive Shock Acceleration (DSA) at supernova remnant blast waves is the favoured production mechanism for the production of the galactic cosmic-rays (CR). This theory, developed since the late 70s (see \\citealt{Drury1983a} for a review), has now both strong theoretical and observational supports. The theoretical grounds of the model lie in the early ideas of Fermi (\\citeyear{Fermi1949a}, \\citeyear{Fermi1954a}): the regular Fermi acceleration mechanism (known as Fermi~I, the stochastic one being known as Fermi~II) can naturally explain the formation of a power-law spectrum by a shock wave -- with a remarkable universal slope whose value $s$ depends solely on the shock compression ratio $r$ (which is always 4 for strong non-relativistic shocks). However, the acceleration process can easily be so efficient that the CR may back-react on the shock dynamics, modifying the acceleration process in a fully non-linear way, and requiring a much more detailed analysis (see \\citealt{Malkov2001c} for a review). Thus the DSA mechanism has still received a lot of attention in the last 20 years, from both a theoretical and a numerical perspective. Analytical works have been mostly limited to the test-particle (linear) regime. The full non-linear time-dependent problem has been mostly investigated through numerical simulations, using several different approaches (see \\citealt{Jones2001a} for a short review). A first class is based on particle methods, from the early Monte-Carlo simulations developed by \\cite{Ellison1984a} to the recent Particle-In-Cells codes (eg \\citealt{Dieckmann2000a}). An alternate approach consists of solving the (Fokker-Planck) transport equation. This has first been done in the \"two fluid\" model (eg \\citealt{Jones1990a}), then dealing with the complete particle distribution function (\\citealt{Falle1987a}, \\citealt{Bell1987a}, \\citealt{Kang1991a}, \\citealt{Duffy1992a}). Most of this work has been aimed at understanding the role of single (isolated) supernovae remnants, although in many contexts CR are likely to experience many shocks, most notably inside superbubbles (\\citealt{Parizot2004a}). In this paper we present a new code for the study of DSA, named \\emph{Marcos} for \\emph{Machine {\\`a} Acc{\\'e}l{\\'e}rer les Rayons COSmiques}\\footnote{the French for \\emph{COSmic-Rays Acceleration Machine}}. In section~\\ref{sec-dsa} we present the basics of the numerical methods implemented in our code, which couples the hydrodynamical evolution of a fluid with the kinetic transport of the CR. In section~\\ref{sec-scales} we present the Adaptive Mesh Refinement (AMR) technique which allows us to resolve the (very) different scales induced by CR diffusion. In section~\\ref{sec-multi} we present parallelization of the code, to be able to study in reasonable wall-clock time the effects of multiple shocks. ", "conclusions": "We have presented a new code aimed at the simulation of time-dependent non-linear diffusive shock acceleration. It is based on the kinetic approach, coupling the hydrodynamical evolution of the plasma with the diffusive transport of the distribution function of the supra-thermal particles. As such it falls under the legacy of the pioneers (\\citealt{Falle1987a}, \\citealt{Duffy1992a}) and of the masters (\\citealt{Kang1991a}, \\citealt{Kang2001a}) of the genre. As the CRASH code it implements an efficient AMR technique to deal with the huge range of space- and time-scales induced by CR diffusion of Bohm-like type. To save even more on computing time we have also parallelized our code in momentum so that we can study acceleration by multiple shocks as fast as acceleration by a single shock. However in many aspects (high-Mach flows, shock tracking, self-consistent injection) our code remains numerically simpler than CRASH -- which can be both a limitation and an advantage. Regarding the physics we note that various mechanisms of importance could be included in the code: self-consistant diffusion coefficient (adding magnetic waves transport), second-order Fermi acceleration (especially between multiple shocks), electrons acceleration (adding radiative losses), CR radiation (hadronic and leptonic)\\ldots We have presented a few tests that show that our code works well, with respect to both the physical accuracy and the numerical efficiency, even in realistic difficult situations. We are now able to investigate in details the various aspects of the DSA mechanism, which 30 years after its early developments still poses some difficulties. In particular we can address the non-linear \\emph{multiple} DSA mechanism, which we believe hasn't received so far all the attention it deserves. Our very first results suggest that the injection fraction plays a crucial role. We intend now to study in more details the situation in superbubbles. %" }, "0710/0710.1108_arXiv.txt": { "abstract": "We study the limits of accuracy for weak lensing maps of dark matter using diffuse 21-cm radiation from the pre-reionization epoch using simulations. We improve on previous ``optimal'' quadratic lensing estimators by using shear and convergence instead of deflection angles. We find that non-Gaussianity provides a limit to the accuracy of weak lensing reconstruction, even if instrumental noise is reduced to zero. The best reconstruction result is equivalent to Gaussian sources with effective independent cell of side length $2.0h^{-1}\\, \\rm Mpc$. Using a source full map from z=10-20, this limiting sensitivity allows mapping of dark matter at a Signal-to-Noise ratio (S/N) greater than 1 out to $l\\lesssim 6000$, which is better than any other proposed technique for large area weak lensing mapping. ", "introduction": "\\label{INTRO} The lens mapping of dark matter is an essential cornerstone of modern precision cosmology. Weak gravitational lensing has developed rapidly over the past years, which allows the measurement of the projected dark matter density along arbitrary lines-of-sight using galaxies as sources. Recently, \\citet{2007PhRvD..76d3510S} have demonstrated the first CMB lensing detection. The goal is now to achieve high precision cosmological measurements through lensing, at better than 1\\% accuracy. Galaxies are plentiful on the sky, but their intrinsic properties are not understood from first principles, and must be measured from the data. Future surveys may map as many as $10^{10}$ source objects. Using galaxies as lensing sources has several potential limits \\citep{2004PhRvD..70f3526H}, including the need to calibrate redshift space distributions and PSF corrections, to be better than the desired accuracy, say 1\\%. This will be challenging for the next generation of experiments. Some sources, such as the CMB, are in principle very clean, since its redshift and statistical properties are well understood. Unfortunately, there is only one 2-D CMB sky with an exponential damping at $l \\gg 1000$, which limits the number of source modes to $\\sim 10^6$. The potential of detecting the 21-cm background from the dark ages will open a new window for cosmological detections. Studying the 21-cm background as high redshifts lensing source, as well as the physics of the 21-cm background itself, provide rich and valuable information to the evolution of universe. The number of modes on the sky is potentially very large, with numbers of $10^{16}$ or more. For this reason, 21-cm lensing has recently attracted attention. However, most of the reconstruction methods are based on a Gaussian assumption \\citep{2004NewA....9..417P,2004NewA....9..173C,2006ApJ...653..922Z, 2006astro.ph.11862B,2007arXiv0706.0849H}. In contrast to CMB lensing, where the Gaussian assumption works well, non-Gaussianity in 21-cm lensing may affect the results. Non-linear gravitational clustering leads to non-Gaussianity, and ultimately to reionization. In this paper, we will address the problem of the lensing of pre-reionization gas. 21-cm emission is similar to CMB: both are diffuse backgrounds. It is natural to apply the techniques used in CMB lensing. \\citet{2002ApJ...574..566H} expand the CMB lensing field in terms of the gravitational potential (or deflection angles), and construct a trispectrum based quadratic estimator of potential with maximum S/N. However, unlike CMB, the 21-cm background has a 3-D distribution and is intrinsically non-Gaussian. A fully 3-D analysis is explored in \\citet{2006ApJ...653..922Z}, where they generalize the 2-D quadratic estimator of CMB lensing \\citep{2002ApJ...574..566H} to the 3-D Optimal Quadratic Deflection Estimator (OQDE). A local estimator was proposed in \\citet{2004NewA....9..417P}, which assumed a power law density power spectrum. In this paper, we will design localized estimators for the lensing fields under the Gaussian assumption, and apply the derived reconstruction technique to Gaussian and non-Gaussian sources. The influence of non-Gaussianity can be measured by comparing the numerical results between the Gaussian sources and non-Gaussian sources. Quadratic lensing reconstruction is a two point function of the lensed brightness temperature field of the 21-cm emission. In the paper, 3-D quadratic estimators are constructed for the convergence ($\\kappa$), as well as the shear ($\\gamma$). Our method recovers the $\\kappa$ and $\\gamma$ directly instead of gravitational potential or deflection angles. Our estimators have in principle the same form as the OQDE, consisting of the covariance of two filtered temperature maps. The OQDE reconstructs the deflection angle, while our estimators reconstruct the kappa and shear fields. Our filtering process can be written as a convolution of the observed fields. As presented in Appendix and section 4, our combined estimator is unbiased, and equally optimal as the OQDE for Gaussian sources, and has better performance for non-Gaussian sources, and recovers three extra (constant) modes. Other authors also developed reconstruction methods from alternative approaches. \\citet{2006astro.ph.11862B} give a estimator for shear. They choose the separate 2-D slices at certain redshift intervals, and then these slices can be treated as independent samples for the same lensing structure. As a result, the information between these slices are lost. \\citet{2004NewA....9..173C} expands the lensed field to a higher order of the gravitational potential, and investigates the higher order correction to the lensed power spectrum. The paper is organized as follows: The basic framework of lensing and the reconstruction method is introduced in $\\S 2$. The numerical methods are presented in $\\S 3$. The results are discussed in $\\S 4$. We conclude in $\\S 5$. ", "conclusions": "\\label{CONC} In this paper, we developed the maximum likelihood estimator for the large-scale structure from the 21-cm emission of the neutral gas before the epoch of re-ionization. The convergence and shears can be constructed independently. To test the effects of non-Gaussianity, we applied our estimators to simulated data. The sources were generated by N-body simulations, because gas is expected to trace the total mass distribution. To investigate the influence of non-Gaussianity, we also use Gaussian sources which have the same power spectrum as the simulated sources. We applied our estimator and the OQDE on both the Gaussian and non-Gaussian sources. Though our estimators are derived in the simplified case of a constant convergence, the noise of our combined estimator of convergence and shear are the same as the OQDE for Gaussian sources. For a finite survey area, three extra constant modes can be recovered. The non-Gaussian nature of the source can increase the error bar by orders of magnitude, depending on the experimental cut off scale. Shear construction is affected less by non-Gaussianity than the convergence field, and the combined estimator with non-Gaussian noise weights is a better choice than reconstructing with the OQDE. S/N can not be boosted infinitely by reducing the experimental noise, and achieves its maximum for a cut off around $k^{\\rm NG}_{\\rm c}\\approx 4h\\,\\Mpc^{-1}$. Below that scale the S/N start to saturate or even decrease. The maximum S/N for non-Gaussian sources is equal to Gaussian sources with $k^{\\rm G}_{\\rm c}\\approx2h\\,\\Mpc^{-1}$, where the power spectrum of source is $\\Delta^2\\approx0.2$ and the side length of the effectively independent cells is $ 2.0 h^{-1}\\,\\rm Mpc$. The maximum S/N is greater than unity for $l\\lesssim 6000$, which makes 21-cm lensing very competitive compared to optical approaches. {\\it Acknowledgments} We thank Oliver Zahn, Chris Hirata, Brice M\\'{e}nard and Mike Kesden for helpful discussions. T.T. Lu thanks Pengjie Zhang, Zhiqi Huang, Hy Trac, and Hugh Merz for help in the early stage of the work." }, "0710/0710.0562_arXiv.txt": { "abstract": "Several studies have correlated observations of impulsive solar activity --- flares and coronal mass ejections (CMEs) --- with the amount of magnetic flux near strong-field polarity inversion lines (PILs) in active regions' photospheric magnetic fields, as measured in line-of-sight (LOS) magnetograms. Practically, this empirical correlation holds promise as a space weather forecasting tool. Scientifically, however, the mechanisms that generate strong gradients in photospheric magnetic fields remain unknown. Hypotheses include: the (1) emergence of highly twisted or kinked flux ropes, which possess strong, opposite-polarity fields in close proximity; (2) emergence of new flux in close proximity to old flux; and (3) flux cancellation driven by photospheric flows acting fields that have already emerged. If such concentrations of flux near strong gradients are formed by emergence, then increases in unsigned flux near strong gradients should be correlated with increases in total unsigned magnetic flux --- a signature of emergence. Here, we analyze time series of MDI line-of-sight (LOS) magnetograms from several dozen active regions, and conclude that increases in unsigned flux near strong gradients tend to occur during emergence, though strong gradients can arise without flux emergence. We acknowledge support from NSF-ATM 04-51438. ", "introduction": "\\label{sec:intro} It has been known for decades that flares and filament eruptions (which form CMEs) originate along polarity inversion lines (PILs) of the radial photospheric magnetic field. In studies using photospheric vector magnetograms, Falconer {\\em et al.}~(2003, 2006) \\nocite{BW_Falconer2003,BW_Falconer2006} reported a strong correlation between active region CME productivity and the total length of PILs with strong potential transverse fields ($>150$ G) and strong gradients in the LOS field (greater than 50 G Mm$^{-1}$). They used a $\\pm$2-day temporal window for correlating magnetogram properties with CMEs. Falconer {\\em et al.}~(2003) \\nocite{BW_Falconer2003} noted that these correlations remained essentially unchanged for ``strong gradient'' thresholds from 25 to 100 G Mm$^{-1}$. Using more than 2500 MDI (LOS) magnetograms, Schrijver (2007) \\nocite{BW_Schrijver2007} found a strong correlation between major (X- and M-class) flares and the total unsigned magnetic flux near (within $\\sim 15$ Mm) strong-field PILs --- defined, in his work, as regions where oppositely signed LOS fields that exceed 150 G lie closer to each other than the instrument's $\\sim$ 2.9 Mm resolution. Schrijver's (2007) effective gradient threshold, {$\\sim$ 100 G Mm$^{-1}$}, is stronger than that used by Falconer {\\em et al.}~(2003, 2006). \\nocite{BW_Falconer2003,BW_Falconer2006} Although these studies were published recently, the association between flares and $\\delta$ sunspots, which posses opposite-sign umbrae within the same penumbra --- and therefore also possess strong-field PILs --- has been well known for some time \\cite{BW_Kunzel1960,BW_Sammis2000}. In particular, $\\beta \\gamma \\delta$ spot groups are most likely to flare \\cite{BW_Sammis2000}. A $\\beta \\gamma$ designation means no obvious north-south PIL is present in an active region \\cite{BW_Zirin1988}. We note that Cui {\\em et al.}~(2006) \\nocite{BW_Cui2006} found that the occurrence of flares is correlated with the maximum magnitude of the horizontal gradient in active region LOS magnetograms --- not just near PILs --- and that the correlation increases strongly for gradients stronger than $\\sim$ 400 G Mm$^{-1}$. One would expect the measures of CME- and flare- productivity developed by both Falconer {\\em et al.}~(2003,2006) \\nocite{BW_Falconer2003,BW_Falconer2006} Schrijver (2007) \\nocite{BW_Schrijver2007} to be larger for larger active regions. Importantly, however, both studies showed that their measures of flux near strong-field PILs is a better predictor of flare productivity than total unsigned magnetic flux. Evidently, more flux is not, by itself, as significant a predictor of flares as more flux near strong-field PILs. These intriguing results naturally raise the question, ``How do strong-field PILs form?'' For brevity, we hereafter refer to strong-field PILs as SPILs. Schrijver (2007) \\nocite{BW_Schrijver2007} contends that large SPILs form primarily, if not solely, by emergence. But he also noted that flux emergence, by itself, does not necessarily lead to the formation of SPILs. Rather, a particular type of magnetic structure must emerge, one containing a long SPIL at its emergence. He suggests such structures are horizontally oriented, filamentary currents. Beyond the ``intact emergence'' scenario presented by Schrijver (2007), \\nocite{BW_Schrijver2007} other mechanisms can generate SPILs. When new flux emerges in close proximity to old flux --- a common occurrence \\cite{BW_Harvey1993} --- SPILs can form along the boundaries between old and new flux systems. Converging motions in flux that has already emerged can also generate SPILs. If the convergence leads to flux cancellation by some mechanism --- emergence of U loops, submergence of inverse-U loops, or reconnective cancellation \\cite{BW_Welsch2006} --- then the total unsigned flux in the neighborhood of the SPIL might decrease as the SPIL forms. We note that, while cancellation in already-emerged fields can occur via flux emergence (from upward moving U-loops), the emergence of a new flux system across the photosphere must increase the total unsigned flux that threads the photosphere. If the emergence of new flux were primarily responsible for SPILs, then a straightforward prediction would be that an increase in total unsigned flux should be correlated with an increase in the amount of unsigned flux near SPILs. Hence, observations showing that increases in the unsigned flux near SPILs frequently occur without a corresponding increase in total unsigned flux would rule out new flux emergence as the sole cause of these strong field gradients. Our goal is to investigate the relationship between increases in the amount of unsigned flux near SPILs with changes in unsigned flux in the active regions containing the SPILs, to determine, if possible, which processes generate SPILs. ", "conclusions": "In Figure \\ref{fig:tres}, we show a scatter plot of changes in $R$ as a function of changes in ${\\cal B}$. The plot does not show the full range in $\\Delta {\\cal B}$, but the $\\Delta R$ for outliers on the horizontal axes are near zero. One striking feature of the plot is its flatness, i.e., that most changes in ${\\cal B}$ are not associated with any change in $R$. In Table \\ref{tab:uno}, we tabulated the data points in each quadrant of this plot. Clearly, increases in $R$, the unsigned flux near SPILs, usually occur simultaneously with increases in the unsigned flux over the entire active region. Increases in $R$ only occur less frequently when flux is decreasing, i.e., during cancellation. \\begin{figure}[!ht] \\includegraphics[width=5.5in]{welb_fig03.eps} \\caption{A scatter plot of changes in $R$ as a function of changes in ${\\cal B}$. Increases in $R$, the unsigned flux near SPILs, usually occur simultaneously with increases in the unsigned flux over the entire active region. Increases in $R$ only occur rarely when flux is decreasing, i.e., during cancellation. For a breakdown of the data points in each quadrant, see Table \\ref{tab:uno}.} \\label{fig:tres} \\end{figure} \\begin{table} \\caption{Breakdown of Flux Changes \\label{tab:uno}} \\begin{tabular}{l|c|c} & $\\Delta {\\cal B} < 0$ & $\\Delta {\\cal B} > 0$ \\\\ \\hline $\\Delta R > 0$ & 215 & 671 \\\\ \\hline $\\Delta R < 0$ & 363 & 371 \\\\ \\end{tabular} \\end{table} We set out to answer the question, ``How do strong-field PILs form?'' We related changes in total, unsigned flux over whole active regions with changes in total, unsigned flux in subwindows of the same active regions --- defined by weighting maps. One might expect, therefore, that these quantities should be correlated, casting doubt about our ability to discrimintate between changes in total flux in active regions and in subwindows. If the two were strongly correlated, the excess of events with $\\Delta R > 0$ and $\\Delta {\\cal B} > 0$ might not be very meaningful. In fact, however, $\\Delta R$ and $\\Delta {\\cal B}$ are poorly correlated: the two have a linear correlation coefficient $r = 0.29$, and a rank-order coefficient of $0.36.$ This suggests that the relationship between increases in $R$ and increases in total, unsigned active region flux is not an artifact of our approach. Nonetheless, our active region sample is not ideally suited to address the origin of SPILs, generally. Our sample was not unbiased with respect to active region morphology; we selected regions with well-defined PILs. In addition, our sample included some decayed active regions that NOAA AR designations. Consequently, we believe that a follow-up study, with a much larger, unbiased sample of active regions, is warranted. With caveats, therefore, our study supports Schrijver's (2007) \\nocite{BW_Schrijver2007} contention that the emergence of new flux creates the strong-field polarity inversion lines that he found to be correlated with flares." }, "0710/0710.5171_arXiv.txt": { "abstract": "We describe a method for computing the biases that systematic signals introduce in parameter estimation using a simple extension of the Fisher matrix formalism. This allows us to calculate the offset of the best fit parameters relative to the fiducial model, in addition to the usual statistical error ellipse. As an application, we study the impact that residual systematics in tomographic weak lensing measurements. In particular we explore three different types of shape measurement systematics: (i) additive systematic with no redshift evolution; (ii) additive systematic with redshift evolution; and (iii) multiplicative systematic. In each case, we consider a wide range of scale dependence and redshift evolution of the systematics signal. For a future DUNE-like full sky survey, we find that, for cases with mild redshift evolution, the variance of the additive systematic signal should be kept below $10^{-7}$ to ensure biases on cosmological parameters that are sub-dominant to the statistical errors. For the multiplicative systematics, which depends on the lensing signal, we find the multiplicative calibration $m_0$ needs to be controlled to an accuracy better than $10^{-3}$. We find that the impact of systematics can be underestimated if their assumes redshift dependence is too simplistic. We provide simple scaling relations to extend these requirements to any survey geometry and discuss the impact of our results for current and future weak lensing surveys. ", "introduction": "\\label{intro} Weak gravitational lensing, or `cosmic shear', is undergoing a phase of rapid expansion \\citep[see][for reviews]{2003ARA&A..41..645R,2006astro.ph.12667M,2003astro.ph.10908H} with many future surveys and instruments being planned (e.g. DUNE\\footnote{http://www.dune-mission.net}, PanSTARRS\\footnote{http://pan-starrs.ifa.hawaii.edu}, DES\\footnote{https://www.darkenergysurvey.org}, SNAP\\footnote{http://snap.lbl.gov} and LSST\\footnote{http://www.lsst.org}). Central to the planning and designing of these instruments is our ability to predict the uncertainties that such measurements will achieve on the cosmological parameters. To this end, the Fisher matrix has become a widely used tool in cosmology for calculating their covariance matrix. However, a limitation of this approach is that it is only able to account for statistical errors, i.e. ones that cause an enlargement of the error bars, and is not well-suited for treatment of systematic errors, which can introduce biases that move the measured central value relative to its true value. One approach that is commonly taken to overcome this limitation is to treat the systematic errors in the same way as statistical errors and to marginalise over possible values. This introduction of nuisance parameters, in addition to the cosmological parameters, causes the error ellipses to expand. A more accurate approach, which we use here, is to directly calculate the bias that the systematic signals will introduce. This bias will tend to offset the central value to the measurements from the true values, as shown in figure \\ref{fig:fig0}. Since the computations needed for this calculation are very similar to those performed in the standard Fisher matrix analysis, extending the current Fisher matrix analysis to include a calculation of bias is relatively straightforward. We apply this formalism to study the impact of residual systematics on tomographic cosmic shear surveys. In particular, we consider systematics arising in the measurement of galaxy shapes after correction of instrumental effects (such as the Point Spread Function). We consider both additive and multiplicative systematics and explore a wide range of scale and redshift dependences. In an earlier work, \\cite{2006MNRAS.366..101H} considered the impact of photometric calibration errors and power law shape systematics using the Fisher matrix formalism. A similar bias formalism was introduced by \\cite{2005APh....23..369H} and applied to theoretical uncertainties in modeling the matter power spectrum with N-body simulations. Our work expands upon these earlier works, by appying the bias formalism to a broad set of shape systematics and by studying the joint impact of systematic and statistical errors in current and future surveys. This paper is organised as follows. In section 2, we describe the formalism that we use to quantify systematic biases. In section 3, we apply our formalism to cosmic shear surveys by exploring the effect of three types of shape measurement systematics: (i) additive with no redshift evolution; (ii) additive with redshift evolution; and (iii) multiplicative. For each type, we consider several possibilities for their scale dependence: (i) log-linear systematics; (ii) systematics that have the same shape as the lensing signal; and (iii) systematics that mimic a small change in the cosmological parameters. In section 4, we study the impact of the systematics in the design of future surveys. Our conclusions are summarised in section 5. ", "conclusions": "\\label{conclusion} In this paper, we have outlined a method for computing the biases that residual systematics introduce. This approach involves a simple extension of the Fisher matrix formalism that is now widely used in cosmology to make error forecasts. As an application, we have used it to study the impact that residual systematic signals will have on future tomographic cosmic shear measurements. Specifically, we have explored three different types of shape systematic signal affecting tomographic shear power spectra: (i) additive systematics with no redshift evolution; (ii) additive systematics with redshift evolution; and (iii) multiplicative systematics. The requirement target is then to have all types of systematics close to zero. This defines a tolerance envelope for the systematics that allows the residual systematics errors in the power spectra to be positive or negative within its limits. It is important to note that it is the $worst$ systematic possible within this limit which drives the requirement, not a marginalised average over all systematics. To this end we have investigated a wide class of possible systematic shapes and used the most constraining ones to set our systematic requirements. In doing this, we have found that, for both the additive and multiplicative parts, it is vital to consider systematics that have positive and negative power spectra $C_\\ell^{sys}$. For instance we see, in the multiplicative case, that investigating only power-law behaviour for its redshift evolution (i.e. $m$ is always positive) can lead to a factor of 5 underestimation of the impact of a systematic within a given tolerance window. From our calculation we are able to set the following requirements on the survey we have considered (a DUNE-like survey covering 20,000 sq. degrees with 35 galaxies per arcmin$^{-2}$ and a median redshift of $z_m=0.9$): \\begin{itemize} \\item For both the additive and the multiplicative signals, the redshift evolution needs to be weak $\\beta < 1.5$, where the errors for a given galaxy scale as $(1+z)^\\beta$. This is not a trivial requirement since the shapes of more distant galaxies are harder to measure since they are smaller and fainter. \\item The power spectrum of the residual additive shear error, that is the part that is not correlated to the lensing signal, must be controlled such that its amplitude is $\\sigma^2_{sys} < 10^{-7}$ (as defined in equation \\ref{eq:var}) \\item The multiplicative part needs to be controlled to a precision of $m_0 <10^{-3}$, where $m_0$ is the shear calibration error. This means that we need to be able calibrate shears to an accuracy of 0.1\\%, which is about one order of magnitude better than the current best measurement methods are able to achieve, as determined by the latest STEP simulations \\citep{2006MNRAS.368.1323H,2007MNRAS.376...13M}. \\end{itemize} These specific requirements apply to our fiducial survey, but we provide scaling relations (Eqs \\ref{eq:scale1} and \\ref{eq:scale2}) which show the requirements for any survey geometry. We have shown that for current survey covering $\\sim 100$ deg$^{2}$ \\citep{2007MNRAS.381..702B}, we need $\\sigma^2_{sys} < 3\\times10^{-6}$ and $m< 0.03$. This level of accuracy is at the limit of the performance of the current best shear measurement methods, as demonstrated by STEP. However, further systematics, such as that arising from PSF calibration and interpolation, are not accounted for by STEP and can dominate the error budget for future surveys. A discussion of the requirements for additive systematics and PSF modeling in the context of present and future surveys will be presented in a later paper (Paulin-Henriksson et al., 2007, in prep). \\newpage \\appendix" }, "0710/0710.3754.txt": { "abstract": "We present arcsecond-scale mid-ir photometry (in the 10.5 $\\mu$m N band and at 24.8 $\\mu$m), and low resolution spectra in the N band ($R\\simeq100$) of a candidate high mass protostellar object (HMPO) in IRAS 18151-1208 and of two HMPO candidates in IRAS 20343+4129, IRS 1 and IRS 3. In addition we present high resolution mid-ir spectra ($R\\simeq80000$) of the two HMPO candidates in IRAS 20343+4129. These data are fitted with simple models to estimate the masses of gas and dust associated with the mid-ir emitting clumps, the column densities of overlying absorbing dust and gas, the luminosities of the HMPO candidates, and the likely spectral type of the HMPO candidate for which [Ne II] 12.8 $\\mu$m\\ emission was detected (IRAS 20343+4129 IRS 3). We suggest that IRAS 18151-1208 is a pre-ultracompact HII region HMPO, IRAS 20343+4129 IRS 1 is an embedded young stellar object with the luminosity of a B3 star, and IRAS 20343+4129 IRS 3 is a B2 ZAMS star that has formed an ultracompact HII region and disrupted its natal envelope. ", "introduction": "Many open questions in high-mass star formation are related to the evolution of circumstellar envelopes, accretion disks, and jets from high-mass protostellar objects (HMPOs). HMPOs are often bright sources in mid-ir continuum, but only in a few recent cases have images suggested specific structures such as disks, jets, or warm outflow cavity walls \\citep{sri05, deb05, deb06,deb07}. Mid-ir ionic lines like [Ne II] and [S IV] have been used to map compact and ultracompact HII regions (UC HII) and photodissociation regions, and to study their structure and excitation \\citep{lac82,oka01,zhu05,kas02,kas06}, but there is a lack of observations of HMPOs. A remarkable feature of high-mass star formation is that HMPOs, defined as actively accreting mass, can begin nuclear fusion and hence also be rapidly evolving massive young stellar objects (MYSOs) that have already formed hypercompact or ultracompact HII regions \\citep{beu07}. This feature raises a possibility of determining the spectral type of an MYSO through the ionic lines excitation, or from the number of ionizing photons required for its observed centimeter continuum flux, separately from estimating its luminosity and spectral type from infrared emission. However there is also the possibility that the ionization is collisionally excited by a jet. It may be possible to distinguish between the two cases, depending on the Doppler velocities, the morphology of ionized gas, and the ratio of the required flux of ionizing photons to total luminosity. Hoping to enlarge the sample of HMPOs that could be studied in detail despite potential limitations, in 2003 we made mid-ir observations on the IRTF of about a third of the survey of 69 HMPO candidates presented by \\citet{sri02} and \\citet{beu02a}. We chose objects from their survey that appeared compact and/or bright in the MSX survey, and found that about 80\\% of them were unresolved or marginally resolved by MIRSI \\citep{deu02} on the IRTF in the broad N band at 10.5 $\\mu$m and in a narrow band filter at 24.8 $\\mu$m. In addition, we obtained MIRSI grism low-resolution spectra (R $\\simeq100$) of ten of them in the N band. In 2006 on Gemini North, we obtained TEXES \\citep{lac02} high-resolution spectra (R $\\simeq80000$) of two HMPO candidates for which we had grism spectra. In this paper we present spectra and photometry of three candidate HMPOs including the two with TEXES spectra: IRAS 18151-1208, IRAS 20343+4129 IRS 1, and IRAS 20343+4129 IRS 3 (hereafter, 18151, 20343 IRS 1, and 20343 IRS 3). We will demonstrate that mid-ir emission from the dust and gas near the HMPO candidates (where it is strongly heated) can be used as a useful probe of temperatures, masses, and luminosities, using simple isothermal clump models, even if each component (envelope, disk, jet, or cavity wall) is not resolved. In combination with observations cited below, we are able to use our new data to infer the nature of each candidate candidate HMPO (e.g. pre-UCHII region HMPO, ZAMS B2 star). The objects chosen are near the centers of complex, large-scale massive molecular outflows mapped by \\citet{beu02b}. 20343 has an apparent large-scale N-S outflow whose red and blue lobes are both extended E-W \\citep{beu02b}, but IRS 1 also has a compact E-W velocity outflow in CO(2-1), while IRS 3 presents an ambiguous situation \\citep{pal06}. All objects show near-ir emission from shocked H$_2$ \\citep{dav04,kum02}. Two of them (18151 and 20343 IRS 3) were observed to have 0.5 and 1.8 mJy 3.6 cm emission, respectively \\citep{car99,sri02}. We observed 20343 IRS 1 and IRS 3 with TEXES on Gemini North based on the 3.6 cm and H$_2$ observations, with the goal of studying the role of ionized gas in them. The 10 $\\mu$m grism spectral shapes of the HMPO candidates fall into three classes: those with deep silicate absorption; those with moderate silicate absorption and an apparent peak at about 8.5 $\\mu$m; and those without an apparent silicate absorption feature but with continuum rising monotonically from short to long wavelengths (Campbell et al. 2007, in preparation). Examples of these shapes can be seen in the UC HII spectra presented by \\citet{fai98}. Since the HMPO candidates were chosen based on IRAS colors similar to UC HII regions \\citep{sri02}, one would expect the HMPO candidates to have similar 10 $\\mu$m spectra. IRAS 20343+4129, was observed with the IRAS LRS and has a silicate absorption feature \\citep{vol91}. The three objects presented here include an example of each of the three classes of grism spectral shapes. Two of the ten candidate HMPO grism spectra show strong [Ne II] lines, IRAS 18247-1147 and 18530+0215. \\citet{sri02} reported relatively strong 3.6 cm fluxes of 47 and 311 mJy, respectively, for them. We did not observe them with TEXES on Gemini in order to focus on the earliest possible HMPO stage associated with ionized gas. Deriving physical information from the continuum spectra is difficult because the actual geometry of the dust distribution is unknown, except that the sizes of the N band and 24.8 $\\mu$m images limit the extent of the emitting structures. Experience with one-dimensional radiative transfer models of spectral energy distributions from UC HII regions \\citep{cam95,cam00,cam04}, and inspection of the new spectra themselves suggest that there are ranges of temperatures in the emitting regions. However, the two-dimensional radiative transfer models of \\citet{deb05b} and \\citet{whi03a,whi03b,whi04} show that orientation of flattened envelopes and outflow cavities dramatically affects the depth of the silicate absorption feature, as does emission and absorption by individual clumps in a clumpy dust cloud in the three-dimensional models of \\citet{ind06}. The recent observations of extended and complex near- and mid-ir emission from HMPOs \\citep{deb05,deb06,sri05} also indicate that one-dimensional models are unrealistic. Nevertheless, a simple three part model will allow us to derive approximate temperatures, column densities, and masses of different dust components. The first component is hot dust that could be (part of) a relatively compact disk, or a clump very near the HMPO candidate; the second is warm dust that could be a more extended (part of a) disk, a clump of dust further out, or perhaps the inner wall of of an outflow cavity; and the third is cold dust in an outer envelope that creates the silicate absorption feature. Deriving information from the ionic lines of an UC HII region is in principle straight-forward. The line fluxes can be corrected for local extinction based on the continuum models discussed above, and then the ratio of [Ne II] to [S IV] fluxes can be used to determine the exciting star's temperature \\citep{lac82,oka01}. The numerical simulation code cloudy (http://www.nublado.org/) can also be used to deduce the star's temperature and luminosity by matching the simulated intensity and spatial extent of free-free, [Ne II], and [S IV] emission to the observations. The star's temperature can then be compared to that of the spectral type deduced from the cm continuum flux. In addition, comparison of Doppler velocities of the ionic lines to those of molecular lines can be used to indicate if the gas is in an UC HII region or a jet. ", "conclusions": "We have presented high resolution mid-ir observations made with MIRSI on the IRTF and TEXES on Gemini North of three HMPO candidates taken from a partial follow up survey of HMPO candidates originally studied at 1.2 mm and radio wavelengths by \\citet{sri02} and \\citet{beu02a}. They are typical for HMPO candidates observed in the follow up survey being compact at 1$\\arcsec$ resolution, having low resolution spectra with strong, moderate, or weak silicate absorption, and with one emitting the [Ne II] line. A simple model of hot dust in emission, warm dust in emission, and cold dust in absorption was developed to fit our 8-13 $\\mu$m low resolution spectra and our 24.8 $\\mu$m photometric points. Even an apparently flat 8-13 $\\mu$m spectrum requires an absorption component if the underlying emission is assumed to be due to hot or warm silicate dust. The temperatures ranged from $\\sim$ 400-1000 K for the hot dust, and $\\sim$ 100-200 K for the warm dust. Using \\citet{dra03a} $R_V$=5.5 model dust properties and gas-to-dust ratio, only small masses of gas and dust in the two emitting components are needed to fit the data. The masses are less than about 1/10 solar mass (often much less) even though these are high or intermediate mass stars, and the mid-ir emission cannot be due the the bulk of the mass in massive accretion disks. The mid-ir is likely to be emitted by the inner walls of outflow cavities and perhaps partly by the surfaces of accretion disks. On the other hand, high column densities, $10^{22} - 10^{23}$ H nucleons cm$^{-2}$, are required for the cold absorption components. These column densities are less than derived from 1.2 mm 11$\\arcsec$ data using \\citet{dra03a} dust, but the discrepancy may be resolved if the slope of the absorption coefficient from far-ir to mm is flattened, as suggested by some observations. Our three component model is not meant to fit either near-ir or far-ir to mm ends of SEDs. Nevertheless, the dust we are modeling in the hot and warm components appears to absorb the bulk of an HMPO's or intermediate mass YSO's photospheric emission, so that the integrated flux of the two model components without application of the cold dust's extinction matches the luminosity as measured including the far-ir by IRAS. The mid-ir measurements together with the model thus give a reasonable way to determine the luminosity for individual HMPOs. The mid-ir emission of IRAS 18151-1208 together with weak 3.6 cm emission and other previous observations suggest that it is an early stage pre-UC HII HMPO whose luminosity is that of a B0 star. TEXES high resolution spectra that cover emission lines from ionized gas can be used to determine the nature of the emission (jet or HII region) and help determine the properties of the underlying star. In the case of IRAS 20343+4129 IRS 1, a lack of [NeII] emission, a well defined compact CO outflow, a moderately strong silicate absorption feature, and a dust model-based luminosity of 1400 L$_\\sun$ imply that it is an intermediate mass YSO whose luminosity is that of a B3 star. For IRAS 20343+4129 IRS 3 observed [Ne II] emission and 3.6 cm free-free emission are consistent with a cloudy model indicating that the object is a B2 ZAMS star. Its weak silicate absorption and small mid-ir based luminosity suggest that it has already disrupted much of its natal envelope. % For IRAS 20343+4129 IRS 3, observed [Ne II] emission and % 3.6 cm free-free emission are consistent with a cloudy model % indicating that the object is a B2 ZAMS star. Its weak % silicate absorption and small mid-ir based luminosity suggest that it has already disrupted % much of its natal envelope. %% If you wish to include an acknowledgments section in your paper, %% separate it off from the body of the text using the" }, "0710/0710.0879_arXiv.txt": { "abstract": "We present the first fully relativistic longterm numerical evolutions of three equal-mass black holes in a system consisting of a third black hole in a close orbit about a black-hole binary. We find that these close-three-black-hole systems have very different merger dynamics from black-hole binaries. In particular, we see complex trajectories, a redistribution of energy that can impart substantial kicks to one of the holes, distinctive waveforms, and suppression of the emitted gravitational radiation. We evolve two such configurations and find very different behaviors. In one configuration the binary is quickly disrupted and the individual holes follow complicated trajectories and merge with the third hole in rapid succession, while in the other, the binary completes a half-orbit before the initial merger of one of the members with the third black hole, and the resulting two-black-hole system forms a highly elliptical, well separated binary that shows no significant inspiral for (at least) the first $t \\sim 1000M$ of evolution. ", "introduction": " ", "conclusions": "" }, "0710/0710.1491_arXiv.txt": { "abstract": "{Mass loss from red giants in old globular clusters affects the horizontal branch (HB) morphology and post-HB stellar evolution including the production of ultraviolet-bright stars, dredge up of nucleosynthesis products and replenishment of the intra-cluster medium. Studies of mass loss in globular clusters also allows one to investigate the metallicity dependence of the mass loss from cool, low-mass stars down to very low metallicities. } {We present an analysis of new VLT/UVES spectra of 47 red giants in the Galactic globular clusters 47 Tuc (NGC 104), NGC 362, $\\omega$ Cen (NGC 5139), NGC 6388, M54 (NGC 6715) and M15 (NGC 7078). The spectra cover the wavelength region 6100--9900 \\AA\\ at a resolving power of $R = 110,000$. Some of these stars are known to exhibit mid-infrared excess emission indicative of circumstellar dust. Our aim is to detect signatures of mass loss, identify the mechanism(s) responsible for such outflows, and measure the mass-loss rates.} {We determine for each star its effective temperature, luminosity, radius and escape velocity. We analyse the H$\\alpha$ and near-infrared calcium triplet lines for evidence of outflows, pulsation and chromospheric activity, and present a simple model for estimating mass-loss rates from the H$\\alpha$ line profile. We compare our results with a variety of other, independent methods.} {We argue that a chromosphere persists in Galactic globular cluster giants and controls the mass-loss rate to late-K/early-M spectral types, where pulsation becomes strong enough to drive shock waves at luminosities above the RGB tip. This transition may be metallicity-dependent. We find mass-loss rates of $\\sim$10$^{-7}$ to $10^{-5}$ M$_{\\odot}$ yr$^{-1}$, largely independent of metallicity.} {} ", "introduction": "Understanding the evolution and mass loss in red giant branch (RGB) and asymptotic giant branch (AGB) stars is of substantial importance in uncovering the history of metal enrichment of the inter-stellar medium (ISM) and subsequent stellar and planetary formation; as well as understanding the evolution of the stars themselves. By observing mass-losing giant branch stars in globular clusters (GCs), we can simultaneously observe a co-eval set of objects within a cluster with similar progenitor mass and identical metallicity; while we can use different clusters to observe the effects of differing ages and masses, but more particularly metallicities, which can range from near-solar metallicity to as little as $\\sim$0.5\\% of solar. Simultaneously, we also glean information about the globular cluster environment itself -- the mass lost from stars forms the intra-cluster medium (ICM). Studies of the ICM show it is removed from the cluster on a timescale much shorter than the time between Galactic Plane crossings (e.g. Boyer et al.\\ 2006; van Loon et al.\\ 2006a). This mass loss exacerbates cluster evaporation, possibly leading to their dispersal on a timescale of $10^{9-10}$ years (van Loon \\& McDonald 2007). Both the RGB and AGB are associated with episodes of strong mass loss, eventually leading to stellar death. Globular cluster stars are thought to lose around 0.2 M$_{\\odot}$ over their time on the RGB (Rood 1973) with up to 0.1 M$_{\\odot}$ being lost in the helium flash at the RGB tip (Fusi-Pecci \\& Renzini 1975, and references therein). This is required to explain the horizontal branch (HB) morphology, which is dependent on the remaining mantle mass. Dupree et al.\\ (2007) have now observationally confirmed mass loss three magnitudes below the RGB tip, thought to be driven by hydrodynamically- or acoustically-dominated chromospheres. Stars reaching the upper slopes of the AGB are even more extreme, losing mass at rates reaching well over $10^{-6}$ M$_{\\odot}$ yr$^{-1}$ (Wood et al.\\ 1983, 1992; van Loon et al.\\ 1999). This is facilitated by thermal pulses -- periodically occurring temporary ignition of a helium shell-burning source inside the star -- and shorter-timescale ($\\sim$10$^{2-3}$ days) radial pulsations at the surface which can lead to shock-wave emission in the extended stellar atmosphere (Fox \\& Wood 1985). In metal-rich AGB stars, dust forming in the outmost parts of these atmospheres is subject to radiation pressure, which forces it and the gas it collides with to be radially accelerated away from the star. It is not yet clear if dust-driving is the dominating effect in RGB and metal-poor AGB winds, and it is possible that Alfv\\'en waves or pressure (P-mode) waves dominate here (Hartmann \\& MacGregor 1980; Pijpers \\& Habing 1989). Indeed, Judge \\& Stencel (1991, hereafter JS91) argue that the mass-loss rate is largely invariant with the driving process and that most cool giants (RGB or AGB) do not have dust as the primary driver of their winds. More recently, Schr\\\"oder \\& Cuntz (2005, hereafter SC05) have suggested an improved Reimers relation (Reimers 1975), by assuming this invariance and using simple chromospheric energy considerations, which is calibrated using the total RGB mass-loss of GC stars. Previous works have attempted to provide evidence for mass loss and calculate mass-loss rates by analysing optical spectral line profiles, either through asymmetries in the line cores (first performed by Deutsch (1956), for the field giant $\\alpha$ Her), line emission (Cohen 1976) or modelling. Line emission has also been used to model a chromosphere with a large bulk motion (Dupree et al.\\ 1984; Mauas et al.\\ 2006). Semi-empirical methods of finding mass-loss rates have also been used to define relationships based on stellar parameters. Mass-loss rates can also be derived from dust masses calculated from IR excess emission (e.g. Origlia et al.\\ 2002 -- hereafter OFFR02; van Loon 2007), though this must assume a value for the ratio of the masses of dust and gas lost. Mira-type long-period variability is only seen in stars with [Fe/H] $\\gtrsim -1$ (Frogel \\& Whitelock 1998), so one would expect that, assuming pulsation assists in driving mass-loss, the mass-loss rate and/or mechanism would vary with metallicity. In this work, we use all of the above techniques to try to identify mass loss and, where possible, estimate mass-loss rates from a sample of 47 such objects from a range of clusters. Ultimately, we aim to assess the r\\^ole of dust, pulsation and chromospheric activity in driving mass loss, and its dependence on metallicity. Our observations are based on new VLT/UVES echelle spectra at very high spectral resolution. We present the spectra and literature photometry in Section 2 of the paper. In Section 3, we use the Kurucz {\\sc atlas9} models (Kurucz 1993) to estimate the effective temperature, metallicity and surface gravity of the stars, then use the near-IR photometry to calculate other physical parameters (luminosity, radius and escape velocity). A full abundance analysis has not been performed, but will become the subject of a subsequent paper. In Section 4, we analyse the H$\\alpha$ and near-IR Ca II triplet lines and from them provide evidence of outflow from the stellar surface. We present a simple model for estimating mass loss in the stellar wind from the H$\\alpha$ profile in Section 5, and use it to calculate mass-loss rates and wind velocities for our sample of stars. This is followed by a discussion in Section 6, in which we also estimate mass-loss rates via other procedures. Our conclusions are presented in Section 7. ", "conclusions": "In this study, we have presented VLT/UVES data of a set of giant branch stars in globular clusters, which are among the highest resolution, high signal-to-noise spectra of their kind. We use these to characterise and quantify the outflow from their atmospheres. We have used Kurucz's {\\sc atlas9} models to determine stellar temperature and, from this, we have calculated basic stellar parameters, which appear typical for red and asymptotic giant branch stars. Many of the stars we have investigated show clear emission in H$\\alpha$, most spectacularly when strong pulsation is present, which is seen to occur at luminosities above the RGB tip. It is also noted that many of the stars show strongly blue-shifted absorption cores, suggesting bulk outflow from the stellar surface. This is also mirrored in some cases in the near-infrared calcium triplet line profiles. In an effort to quantify this outflow, we have constructed a simple model for the warmer and more metal-poor stars. We have calculated terminal velocities for the winds which are of order 10 km s$^{-1}$ -- much lower than the escape velocity of the stars -- and mass-loss rates of $\\approx$ 10$^{-7}$ to 10$^{-5}$ M$_\\odot$ yr$^{-1}$, which lie well above theoretical expectations, but are consistent with the mass-loss rates derived from IR emission from circumstellar dust. These models also suggest that an emissive shell exists close to the stellar surface in many stars, with the wind probably being largely isothermal beyond it. Outflow velocities of the emission are of order 40 km s$^{-1}$, and mass-loss rates derived from this emission are similar to the ones we obtain from modelling the absorption profile. In the most extreme cases we may have either over-estimated the mass-loss rate from our model, or under-estimated the temperature of the emitting shell. Mass-loss rates correlate weakly with luminosity, but stars showing strong IR excesses (linked with dust production) do not necessarily exhibit higher gas mass-loss rates. We find no correlation between mass-loss rate and metallicity. We suggest that the emission and mass-loss in early-type ($\\lsim$ K3--K5) giants are dominated by a warm chromospheric region, as suggested by some studies, though late-type ($\\gsim$ M3) giants have mass loss dominated by pulsation. It seems likely that the spectral type of the transition between the two regimes is metallicity-dependent, occurring at later spectral types for higher metallicities. The outward velocity of the warm emissive shells associated with the chromospheres is similar to the escape velocity at their anticipated radius (2 to 3 R$_\\ast$), allowing the chromosphere to be the sole driver of mass loss in these stars." }, "0710/0710.5421_arXiv.txt": { "abstract": "\\footnotesize\\ This paper\\footnote{Published in 1999 in the book {\\it Solar Polarization}, edited by K.N. Nagendra \\& J.O. Stenflo. Kluwer Academic Publishers, 1999. (Astrophysics and Space Science Library ; Vol. 243), p. 73-96} addresses the modelling issue of the second solar spectrum. This is the name given to the linearly polarized spectrum which can be observed close to the solar limb using spectro-polarimeters of high polarimetric sensitivity (Stenflo and Keller, 1997). The second solar spectrum is due to scattering processes and offers a rich diagnostic potential for exploring solar magnetic fields via the Hanle effect. However, it is full of mysterious spectral features that cannot be understood with simplified polarization transfer theories, thus suggesting that the underlying scattering physics is more complex than previously thought. In this paper we argue that understanding the second solar spectrum requires the consideration of scattering processes in {\\it multilevel} atomic models, taking fully into account the transfer of {\\it atomic polarization} among {\\it all} the levels involved. To give support to this statement, we begin by pointing out the drastically different predictions, given by the standard resonance line polarization theory, with respect to the emergent polarization in three {\\it different} line transitions. This standard theory neglects the atomic polarization of the {\\it lower} level of the line transition under consideration, i.e. it assumes that there are no population imbalances among the lower-level sublevels. The density matrix polarization transfer theory is then applied to formulate the scattering polarization problem taking properly into account atomic polarization in both the upper and the lower line levels. The consideration of lower-level atomic polarization leads to coupled {\\it non-linear} and {\\it non-local} sets of equations, even for the two-level model atom case considered in this paper. The {\\it unknowns} of these equations are the irreducible tensor components of the atomic density matrix whose {\\it self-consistent} values have first to be obtained to be able to calculate the emergent Stokes profiles. To solve this non-LTE problem of the $2^{nd}$ kind we present some iterative methods that are very suitable for developing a general multilevel scattering polarization code. With these numerical methods some model calculations are performed in order to demonstrate that the inclusion of lower-level atomic polarization leads to similar emergent linear polarization signals in such three {\\it different} line transitions, as some observations show. After pointing out that the ``Na solar paradox'' (Landi Degl'Innocenti, 1998) might admit, in principle, a {\\it multilevel} solution, the paper ends establishing a new solar paradox: ``the Mg solar paradox'', for which no {\\it multilevel} solution seems to be possible. This new result demonstrates that there indeed exists {\\it ground} and {\\it metastable}-level atomic polarization in the solar chromosphere and it suggests that the solution to these ``solar paradoxes'' is to be found by carefully revising our current ideas about the chromospheric magnetic field. ", "introduction": "The ``second solar spectrum'' is a term adopted by Stenflo and Keller (1997) to refer to the remarkable observational discovery that the whole solar spectrum is linearly polarized, when observations are made close to the solar limb using novel spectro-polarimeters that allow the detection of very low amplitude polarization signals ($\\sim10^{-5}$ in the degree of linear polarization). This term is certainly adequate because, as pointed out by these authors, the linearly polarized solar spectrum has a structural richness that often exceeds that of the ordinary intensity spectrum. It is indeed as if the Sun has presented us with an entirely new spectrum to explore. In fact, the second solar spectrum contains a wealth of ``inexplicable'' spectral features which are the signature of physical processes that are presently challenging physicists working in the field. This paper deals with the modelling issue of the second solar spectrum. This modelling requires the solution of a formidable numerical problem that is considered as one of the most challenging tasks of solar and stellar polarimetry. It consists in calculating, for {\\it multilevel} atomic models, the excitation and ionization states of chemical species of given abundance that are {\\it consistent} with the polarization properties of the radiation field produced by such species in any medium of given temperature, density, macroscopic velocity and magnetic field vector. Once this self-consistent atomic excitation is known along the line of sight, it is straightforward to solve the transfer equations for the Stokes parameters in order to calculate the emergent polarization profiles that are to be compared with spectro-polarimetric observations. It is indeed a very complex problem because, in the polarization transfer case, one has to take into account that each level of total angular momentum value X has associated with it (2X+1) sublevels, with ${\\vec {\\rm X}}={\\vec {\\rm J}}= {\\vec {\\rm L}}+{\\vec {\\rm S}}$ if fine structure due to the spin-orbit LS coupling is assumed, or with ${\\vec {\\rm X}}={\\vec {\\rm F}}={\\vec {\\rm J}}+{\\vec {\\rm I}}$ if hyperfine structure due to the nuclear angular momentum $\\vec {\\rm I}$ is taken into account. The populations of these sublevels are sensitive to the polarization and anisotropy state of the radiation field at each point within the medium. Moreover, quantum interferences (or coherences) among the sublevels themselves may also appear, coherences that depend on the energy separation between the levels and on their splitting. These coherences must also be properly quantified to fully specify the excitation state. An additional complication stems from the fact that, in the polarized case, instead of the standard radiative transfer (RT) equation for the specific intensity, one has to solve, in general, a vectorial transfer equation for the four Stokes parameters. Obviously, accounting for this complexity requires working within the framework of a robust theory for the generation and transfer of polarized radiation. I believe that the most rigorous (and suitable) theoretical framework to work with is that developed by Landi Degl'Innocenti (1983), which is based on the irreducible tensor components ($\\rho^K_Q$) of the atomic density matrix (see also Bommier and Sahal-Br\\'echot, 1978). According to this formalism, to each level of angular momentum value X, there correspond $\\rm (2X+1)^{2}$ density-matrix elements. These $\\rho^K_Q$-elements contain information about the populations of the atomic sublevels and about the coherences among them. For instance, for a level with total angular momentum J=1 we have that $\\rho_0^0=(N_1+N_0+N_{-1})/{\\sqrt{3}}$, $\\rho^1_0=(N_1\\,-\\,N_{-1})/{\\sqrt{2}}$ and $\\rho_0^2=(N_1-2N_0+N_{-1})/{\\sqrt{6}}$, where $N_{i}$ (with $i$=1, 0 and -1) are the populations of the three magnetic sublevels. Thus, $\\rho_0^0$ gives the total population of the level, while $\\rho^1_0$ (the {\\it orientation} coefficient) and $\\rho^2_0$ (the {\\it alignment} coefficient) inform us about the {\\it population differences} among the sublevels. Finally, $\\rho^K_Q$-terms with $Q\\neq0$ account for the coherences between Zeeman sublevels whose magnetic quantum numbers differ by $Q$. One of the reasons that explain why the density matrix formalism is so suitable for dealing with the generation and transfer of polarized radiation is that, through an emission process, polarization in spectral lines can originate locally either by the splitting of the atomic levels (splitting that can in turn be due either to the Zeeman or the Stark effect) or by the presence of {\\it population differences} and/or {\\it coherences} among the sublevels. Thus, the $\\rho^K_Q$ elements whose self-consistent values are sought at each spatial grid-point, indeed provide the most suitable way of quantifying, at the atomic level, the information that we need to be able to calculate all the ``sources'' and ``sinks'' of polarization within the medium under consideration. The main criticism of this QED theory is that it is based on the approximation of complete frequency redistribution (CRD), i.e. on an assumption that is not adequate for modelling the polarization of several diagnostically important spectral lines. Fortunately, some very recent work has successfully started to incorporate the effects of partial redistribution (PRD) into the framework of the density matrix formalism (see Landi Degl'Innocenti {\\it et al.}, 1997; Bommier, 1997 a,b). These recent efforts to generalize the density matrix theory to PRD are truly important and should be continued because the CRD theory cannot be applied when coherences between non-degenerate levels are present unless the spectrum of the radiation is flat across a frequency interval $\\Delta\\nu$ centred on the line frequency and larger than the frequency separation between the two levels connected by the coherence (see Landi Degl'Innocenti {\\it et al.}, 1997). The problem of finding the self-consistent values of the irreducible tensor components of the atomic density matrix has been called ``the non-LTE problem of the $2^{nd}$ kind'' (see Landi Degl'Innocenti, 1987). It requires solving jointly the statistical equilibrium (SE) equations for the density matrix elements associated with each level of the assumed atomic model and the Stokes-vector transfer equations for all the radiative transitions involved. This terminology also seems appropriate because, as I shall try to argue below, a better understanding of many of the ``inexplicable'' spectral features of the second solar spectrum can only be achieved by carefully formulating and solving {\\it multilevel} non-LTE problems of the $2^{nd}$ kind, taking fully into account the possibility of {\\it atomic polarization} at {\\it all} the levels of the chosen multilevel atomic model and including the depolarizing role of elastic collisions and magnetic fields. The word ``inexplicable'' in the preceding paragraphs refers to the impossibility of explaining some of the spectral features of the second solar spectrum by means of theories based on the approximation of neglecting atomic polarization in the {\\it lower} level of the line transition under consideration, i.e. theories based on the assumption that there are no significant differences in the populations of the lower-level sublevels or coherences among them. One example of a mysterious feature of the second solar spectrum that has triggered some recent theoretical work (see Trujillo Bueno and Landi Degl'Innocenti, 1997; Landi Degl'Innocenti, 1998) is the linear polarization pattern observed around the Na I ${\\rm D}_2$ and ${\\rm D}_1$ lines. In fact, in a recent letter in {\\it Nature} that demonstrates the robustness of the density matrix polarization transfer theory including partial frequency redistribution effects, Landi Degl'Innocenti (1998) concluded that the observed Na ${\\rm D}_2$ and ${\\rm D}_1$ linear polarization pattern can be explained by assuming the presence of an amount of {\\it ground}-level atomic polarization as important as (and indeed slightly larger than!) that of the upper level. However, because very small {\\it non-vertical} magnetic fields (and/or elastic collisions) destroy the ground-level atomic polarization of Na, Landi Degl'Innocenti (1998) was forced to rule out in the solar chromosphere both elastic collisions and the existence of turbulent magnetic fields and of horizontal, canopy-like fields stronger than $\\sim0.01$ gauss. This has led to an exciting apparent paradox in solar physics because there are observational evidence for both turbulent fields of the order of 10 gauss and canopy-like horizontal fields (see Jones, 1984; Solanki and Steiner, 1990; Faurobert-Scholl, 1992; Bianda {\\it et. al}, 1998; Stenflo {\\it et. al}, 1998). The argument in favour of the simplifying approximation of neglecting lower-level polarization is that the lower level of a line transition is generally long-lived, and that it must thus have plenty of time to be depolarized by elastic collisions and/or weak magnetic fields (Stenflo, 1994; 1997). However, this is expected to be the case concerning only the {\\it ground} and {\\it metastable} levels of atomic systems. It is indeed a major simplifying approximation because it implies that the scattered radiation can be expressed linearly in terms of the incident radiation if stimulated emission processes are also neglected (see Landi Degl'Innocenti, 1984). In other words, the approximation of neglecting lower-level atomic polarization allows the study of scattering line polarization problems in terms of phase matrices that are decoupled from the SE equations. However, as we shall show below, the consideration of lower-level atomic polarization leads to a coupled system of {\\it non-linear} equations, even for the simplest case of a two-level model atom. My motivation for writing this paper is two-fold. Firstly, I would like to provide more arguments concerning the idea (already put forward in the paper by Trujillo Bueno and Landi Degl'Innocenti, 1997) that lower-level atomic polarization is an essential physical ingredient for understanding the second solar spectrum. Secondly, I aim to demonstrate that, contrary to some general beliefs, the density-matrix polarization transfer theory does have a suitable form for practical applications. To these ends, I will consider here three types of line transitions in the solar atmosphere: ($a$) lines with ${\\rm J}_{l}=0$ and ${\\rm J}_u=1$, ($b$) lines with ${\\rm J}_{l}=1$ and ${\\rm J}_u=0$, and ($c$) lines with ${\\rm J}_{l}=1$ and ${\\rm J}_u=1$. Section 2 summarizes the predictions of the standard theory, which neglects lower-level atomic polarization. Section 3 is dedicated to outlining the formulation of the scattering line polarization problem taking into account atomic polarization in both the upper and the lower levels of such line transitions. Here I will show the self-consistent solution for the density matrix elements and the corresponding emergent fractional linear polarization that results from this more correct treatment. Section 4 discusses model calculations including the depolarizing role of elastic collisions. As we shall see, my two-level atom scattering line polarization calculations suggest that, if we aim at understanding the second solar spectrum, we need to consider scattering processes in {\\it multilevel} atomic models, taking fully into account the transfer of atomic polarization among all the atomic levels involved. Section 5 argues that, in principle, the above-mentioned ``Na solar paradox'' might admit a {\\it multilevel} solution. However, Section 6 shows that the observed fractional linear polarization in the Mg $b$-lines can only be explained by invoking the presence of atomic polarization in the lower {\\it metastable} levels of the Mg $b_1$ and $b_2$ lines, thus establishing a new paradox: the ``Mg solar paradox''. Finally, Section 7 gives some concluding remarks after pointing out that there is no multilevel solution to this ``Mg solar paradox''. The Appendix is dedicated to a brief description of some iterative methods for the solution of non-LTE problems of the $2^{nd}$ kind, which I consider as ``the road to be taken'' for the development of a general multilevel scattering line polarization code. ", "conclusions": "There is a crucial difference between the Na and Mg ``solar paradoxes''. For Na there was still the chance that the multilevel scenario outlined above might help to solve it. However, there is no similar multilevel solution for the ``Mg solar paradox'', which leads to the conclusion that there must indeed exist {\\it ground}-level and {\\it metastable}-level atomic polarization in the solar chromosphere. The three Mg $b$ lines share the same upper level and what cannot be understood by means of the standard transfer theory (which neglects lower-level polarization) is that the observed linear polarization in these three Mg $b$ lines turns out to be similar. Thus, even if one chooses multilevel atomic models for Mg to calculate the self-consistent values of the density-matrix elements, and then solves the Stokes transfer equations to get the emergent fractional linear polarization, but neglecting the dichroism contribution that comes from the atomic alignment of the lower {\\it metastable} levels, one would find again that the ensuing prediction is wrong. The only way I see for increasing the emitted polarization in the Mg $b_1$ and $b_2$ lines, so as to bring it to the same level of that corresponding to the Mg $b_4$ line (that has ${\\rm J}_l=0$), is via the {\\it dichroism} contribution (i.e. the term $-\\eta_{\\rm Q}{\\rm I}$ of Eq. 8). As discussed in Section 3, this {\\it dichroism} can only arise if the magnetic sublevels of the ${\\rm J}_l=2$ and ${\\rm J}_l=1$ {\\it metastable} lower-levels of the $b_1$ and $b_2$ lines are {\\it unequally populated}. The same explanation can be given for other groups of lines belonging to other atoms, and arising from a similar multiplet ($^3{\\rm P}^{\\rm o}\\,-\\,^3{\\rm S}$), with the ensuing appearance of extra ``paradoxes'' for other chemical elements. As we have seen, the anisotropic illumination of the atoms in a stellar atmosphere can lead to large population imbalances among the lower-level sublevels of many spectral lines. The modelling of the second solar spectrum requires the reliable calculation of the atomic polarization of the lower and upper levels corresponding to the line transitions of interest. To this end, it is crucial to be able to consider {\\it multilevel} atomic models. This goal can presently be achieved by formulating the problems of interest within the framework of the density matrix polarization transfer theory (see Landi Degl'Innocenti, 1983) and by numerically solving the ensuing non-linear and non-local equations with the iterative methods presented in the Appendix. Only when we know the {\\it self-consistent} values of the alignment coefficients of the Na and Mg atomic levels in several solar atmospheric models shall we be able to figure out a possible solution to such ``solar paradoxes''. Obviously, we are facing a complex problem here, from the observational, theoretical and modelling viewpoints. But it is a highly interesting one, not only because of the fascinating physics that it involves, but mainly because in trying to clarify it we may learn something new about the sun." }, "0710/0710.0668_arXiv.txt": { "abstract": "Molecular gas has been searched for and found in unexpectedly large quantities in some collisional debris of interacting galaxies: HI-rich tidal tails, bridges and collisional rings. It was so far observed through millimeter observations of the CO line and detected towards or near regions of star-formation associated to dense condensations of the atomic hydrogen. The discovery of cool H$_{2}$ at distances greater than 50 kpc from the parent (colliding) galaxies, whereas the external disk of spirals is generally considered to be CO--poor, raised question on its origin and favored the hypothesis of a local production out of collapsed HI clouds. However recent observations of a diffuse CO component along tidal debris have challenged this idea. Another recent puzzle is the measurement in the collisional debris of two interacting systems and four recycled objects of a missing mass, whereas no dark matter is expected there. One debated interpretation is that this unseen component is cold, ``invisible\" molecular gas initially present in the disk of spirals. ", "introduction": "The collisional debris addressed here refer to all the material that is expelled into the intergalactic medium during galaxy--galaxy interactions. This is the result of either the tidal forces that shape bridges and tails, of direct high-speed impacts at the origin of rings or head-on collisions forming systems similar to the so--called ``Taffy\" galaxies (see Struck 1999 for a review). Collisional debris may consist of old and young stars, gas clouds and dust with a relative proportion that depends very much on the type of interactions and properties of the parent galaxies. Tidal tails with a prominent old stellar population will be common in slow encounters between late-type galaxies -- see the famous Antennae galaxies. Purely stellar tails may exist, when the gaseous counterpart has for some reason been displaced and is now offset (Mihos 2001). More commonly, the gaseous tail extends further than the stellar one, because it was originally more extended in the disk of its parent galaxies. Collisional rings are particularly gas-rich, being embedded in a faint stellar halo. Finally, long, purely gaseous, tidal tails may form during high-speed collisions, as recently shown by Duc \\& Bournaud (2007). Finally, if the gas in the collisional debris is dense enough, it collapses and locally form a new generation of stars. Whereas stellar debris have been known for a long time -- they are the peculiarities in the optical catalog of Peculiar Galaxies by Arp (1966) --, their gaseous, HI, counterparts have been systematically studied from less than twenty years (e.g., Hibbard \\& van Gorkom 1996). This was possible thanks to interferometers like the Very Large Array which offer a very large field of view. In situ star formation in collisional debris has been discovered through several ways, including broad-band optical (e.g. Schombert et al. 1990) and UV (e.g., Neff et al. 2005; Boquien et al., 2007) imaging, narrow-band H$\\alpha$ images (e.g. Iglesias-P{\\'a}ramo \\& V{\\'{\\i}}lchez 2001) as well as mid-infrared photometry. If ongoing star-formation is present in collisional debris, molecular gas should be present as well. This paper reviews the various attempts that have been done so far to detect it. ", "conclusions": "Our detailed, multi-wavelength, studies of collisional debris, especially its molecular gas content, raise a number of questions. \\begin{itemize} \\item Where does the cool H$_{2}$ detected in quantities in tidal tails come from? Locally, out of tidally expelled, dense, HI clouds, as suggested by the kinematical data, and/or from the disk of the parent galaxy, as indicated by the detection of more diffuse CO emission outside the main centers of star-formation? The answer to this question will require to carry-out complete maps of the molecular gas along collisional debris. So far most of the CO detections result from pointed observations, biased towards regions of high HI column density. \\item How does the star-formation efficiency (the ratio of the star formation rate to the molecular mass) in collisional debris compare with that of individual star-forming regions of spiral disks and classical dwarf galaxies? A first rough comparison is presented in Braine et al. (2001) but should now be updated. The SFE should in particular be determined in an homogeneous way, a difficult task for objects of different sizes and environments. \\item Is really the missing mass found in collisional debris and hence in spiral disks very cold H$_{2}$? First of all, the method used to determine the dynamical mass could be challenged. The objects formed in the debris are, spatially, barely resolved and their inclination is not directly determined; however the uncertainties of the method have been well assessed by numerical simulations. Besides, the conversion factor used to derive the mass of luminous cool H$_{2}$ from the CO line intensity is not very constrained; in the relative high-Z environment of TDGs, it should however be not so different than the assumed galactic value. If real, the mismatch between the dynamical and the luminous mass, estimated to a factor of 2--3 in now 4 different objets, can also be interpreted in a totally different manner. Cold streams of non-baryonic dark matter in disks is a theoretical possibility. Gentile et al. (2007) and Milgrom (2007) were able to reproduce the observed velocity curves of the recycled objects around NGC~5291 within the MOND framework. On the other hand, new simulations including conventional gravity, usual cosmological Dark-Matter halos and the putative cold component of H$_{2}$ in spiral disks (which was not implemented in the numerical modes of Bournaud et al. 2007) seem to also give a good match with the data (Revaz et al., in preparation). \\end{itemize} Although being just ``debris\", tidal tails, rings and more in general all the material sent out of galaxies during their merging history, especially the molecular gas, may tell about some fundamental aspects of astrophysics." }, "0710/0710.1897_arXiv.txt": { "abstract": "Inspiral signals from binary compact objects (black holes and neutron stars) are primary targets of the ongoing searches by ground-based gravitational-wave interferometers (LIGO, Virgo, GEO-600 and TAMA-300). We present parameter-estimation simulations for inspirals of black-hole--neutron-star binaries using Markov-chain Monte-Carlo methods. For the first time, we have both estimated the parameters of a binary inspiral source with a spinning component and determined the accuracy of the parameter estimation, for simulated observations with ground-based gravitational-wave detectors. We demonstrate that we can obtain the distance, sky position, and binary orientation at a higher accuracy than previously suggested in the literature. For an observation of an inspiral with sufficient spin and two or three detectors we find an accuracy in the determination of the sky position of typically a few tens of square degrees. ", "introduction": "\\label{sec:intro} Binary systems with compact objects --- neutron stars (NS) and black holes (BH) --- in the mass range $\\sim 1\\,\\Ms - 100\\,\\Ms$ are among the most likely sources of gravitational waves (GWs) for ground-based laser interferometers currently in operation (Cutler \\& Thorne, 2002): LIGO (Barish \\& Weiss 1999), Virgo (Arcese et al.\\ 2004), GEO-600 (Willke et al.\\ 2004) and TAMA-300 (Takahashi et al.\\ 2004). Merger-rate estimates are quite uncertain and for BH-NS binaries current detection-rate estimates reach as high as 0.1\\,yr$^{-1}$ (\\emph{e.g.} O'Shaughnessy et al.\\ 2008) for first-generation instruments. Upgrades to Enhanced LIGO/Virgo (2008--2009) and Advanced LIGO/Virgo (2011--2014) are expected to increase detection rates by factors of about $\\sim 8$ and $10^3$, respectively. The measurement of astrophysical source properties holds major promise for improving our physical understanding and requires reliable methods for parameter estimation. This is a challenging problem because of the large number of parameters ($> 10$) and the presence of strong correlations among them, leading to a highly-structured parameter space. In the case of high mass ratio binaries (\\emph{e.g.} BH-NS), these issues are amplified for significant spin magnitudes and large spin misalignments (Apostolatos et al.\\ 1994; Grandcl{\\'e}ment et al.\\ 2003; Buonanno et al.\\ 2003). However, the presence of spins benefits parameter estimation through the signal modulations, although still presenting us with a considerable computational challenge. This has been highlighted in the context of LISA observations (see Vecchio 2004; Lang \\& Hughes 2006) but no study has been devoted so far to ground-based observations. In this \\emph{Letter} we examine for the first time the potential for parameter estimation of spinning binary inspirals with ground-based interferometers, including twelve physical parameters. Earlier studies (\\emph{e.g.} Cutler \\& Flanagan 1994, Poisson \\& Will 1995, Van den Broeck \\& Sengupta 2007) have estimated the theoretical accuracy with which some of these parameters should be measured, without determining the parameters themselves. Also, (R\\\"{o}ver et al.\\ 2006, 2007) have explored parameter estimation for non-spinning binaries. We focus on BH-NS binaries where spin effects are strongest (Apostolatos et al.\\ 1994), while at the same time we are justified to ignore the NS spin. We employ a newly developed Markov-chain Monte-Carlo (MCMC) algorithm (van der Sluys et al.\\ 2008) applied on spinning inspiral signals injected into synthetic ground-based noise and we derive posterior probability-density functions (PDFs) of all twelve signal parameters. We show that although sky position is degenerate when using two detectors, we can still determine the mass and spin parameters to reasonable accuracy. With three detectors, the sky position and binary orientation can be fully resolved. We show that our accuracies are good enough to associate an inspiral event with an electromagnetic detection, such as a short gamma-ray burst (\\emph{e.g.} Nakar 2007). ", "conclusions": "\\label{sec:concl} We have explored for the first time the parameter estimation of all physical parameters --- including masses, spin, distance, sky location and binary orientation --- on ground-based gravitational-wave observations of binary inspirals with spinning compact objects. We show that for two detectors and sufficient spin ($a_\\mathrm{spin}\\geq0.5$) or for three detectors, the obtained accuracy in sky position, distance and time of coalescence is good enough to allow the identification of electromagnetic counterparts of compact-binary mergers, \\emph{e.g.} short gamma-ray bursts (Nakar 2007). A direct measurement of mass, spin, distance and orientation can be obtained from inspiral GWs, which is notoriously difficult for electromagnetic observations. The analysis presented here is the first step of a more detailed study that we are currently carrying out, exploring a much larger parameter space, developing techniques to reduce the computational cost of these simulations, and testing the methods with actual LIGO data. The waveform model used here, though adequate for exploratory studies, is not sufficiently accurate for the analysis of real detections, and we are finalising the implementation of a more realistic waveform. Simulations with this improved waveform may also shed light on the degeneracy between mass and spin parameters discussed in Sect.\\,\\ref{sec:results}, and may improve the accuracy of our parameter estimation appreciably (\\emph{e.g.} Van den Broeck \\& Sengupta 2007). Finally, we intend to further develop our Bayesian approach into one of the standard tools that can be included in the analysis pipeline used for the processing of the `science data' collected by ground-based laser interferometers." }, "0710/0710.3776_arXiv.txt": { "abstract": "% The opacity of a spiral disk due to dust absorption influences every measurement we make of it in the UV and optical. Two separate techniques directly measure the total absorption by dust in the disk: calibrated distant galaxy counts and overlapping galaxy pairs. The main results from both so far are a semi-transparent disk with more opaque arms, and a relation between surface brightness and disk opacity. In the Spitzer era, SED models of spiral disks add a new perspective on the role of dust in spiral disks. Combined with the overall opacity from galaxy counts, they yield a typical optical depth of the dusty ISM clouds: 0.4 that implies a size of $\\sim$ 60 pc. Work on galaxy counts is currently ongoing on the ACS fields of M51, M101 and M81. Occulting galaxies offer the possibility of probing the history of disk opacity from higher redshift pairs. Evolution in disk opacity could influence distance measurements (SN1a, Tully-Fisher relation). Here, we present first results from spectroscopically selected occulting pairs in the SDSS. The redshift range for this sample is limited, but does offer a first insight into disk opacity evolution as well as a reference for higher redshift measurements. ", "introduction": "The number of distant galaxies seen in an {\\em HST} image through the face-on foreground spiral is a direct indication of its opacity, after proper calibration using artificial galaxy counts \\citep[synthetic field method (SFM)][]{Gonzalez98,Holwerda05a}. We have obtained calibrated galaxy counts for a sample of 32 deep {\\em HST/WFPC2} fields. The main results from the disk opacity study are: (1) most of the disks are semitransparent \\citep[][Figure \\ref{f:ra}]{Holwerda05b}, (2) arms are more opaque \\citep{Holwerda05b}, (3) as are brighter sections of the disk \\citep{Holwerda05d}. We did not find a relation between HI profiles and disk opacity but this can be better explored on a single disk \\citep{Holwerda05c}. The optimal distance of the foreground disk is $\\sim$ 10 Mpc, the compromise between crowding effects and solid angle constraints \\citep{Gonzalez03,Holwerda05e}. \\begin{SCfigure} \\centering \\includegraphics[width=0.5\\textwidth]{holwerda2.eps}% \\caption{The relation between the optical depth from the SED model [$\\tau_m$(SED)], and the observed optical depth from the number of distant galaxies [$\\tau$(SFM)]. Model C from \\protect\\cite{Natta84} is fit through these. Cloud optical depth is 0.4, and more than a single cloud along the line-of-sight is needed, especially for optically thick disks. Figure from \\protect\\cite{Holwerda07a}.} \\label{f:tc} \\end{SCfigure} There is an overlap of 12 galaxies with {\\em SINGS}; we compared the disk opacity from number of distant galaxies to the dust surface density obtained from an SED model \\citep{Draine07a}. Expressed as optical depths, the relation between the {\\it observed} optical depth ($\\tau$) and the optical depth inferred from the SED dust mass ($\\tau_m$) is solely a function of the typical cloud optical depth ($\\tau_c$): $\\tau/\\tau_m = (1-e^{-\\tau_c})/\\tau_c$ \\citep{Natta84}. Figure \\ref{f:tc} shows the values for $\\tau$ and $\\tau_m$ and the fit of $\\tau_c$=0.4. This implies that most of the disks are made up of small, optically thin, cold ISM structures with more than one cloud along the line-of-sight, especially in optically thick disks \\citep{Holwerda07a}. From their distribution in the E[I-L] color map, it becomes clear that the number of distant galaxies does not drop exclusively as a result of grand spiral opaque structures but that the unresolved dusty ISM disk is equally important \\citep{Holwerda07b}. \\begin{figure} \\centering \\includegraphics[width=0.6\\textwidth]{holwerda3.ps}% \\caption{\\label{f:raz}Radial opacity profile for the local (lower), z=0-0.1 (middle) and z=0.1-0.2 (top) spirals from occulting pairs. Open circles are disk sections for the local pairs. Figure from \\protect\\cite{Holwerda07c}.} \\label{f:tt} \\end{figure} ", "conclusions": "" }, "0710/0710.4360_arXiv.txt": { "abstract": "\\footnote{To appear in proceedings of the Puerto Vallarta Conference on ``New Quests in Stellar Astrophysics II: Ultraviolet Properties of Evolved Stellar Populations'' eds. M. Chavez, E. Bertone, D. Rosa-Gonzalez \\& L. H. Rodriguez-Merino, Springer, ASSP series. Presented as part of a Ph. D. thesis in the Departamento de F\\'\\i sica, of the Universidad de Guadalajara, M\\'exico.} Ultracompact (UC) HII regions with Extended Emission (EE) are classical UC~HII regions associated with much larger ($>$1$'$) structures of ionized gas. The efforts to investigate, detect, and understand if the EE is physically related to the UC emission are few. If they are related, our understanding of UC~HII regions may be affected (e.g., in the estimation of ionizing UV photons). Here we present a brief overview of UC~HII regions with EE (UC~HII+EE) including our most recent effort aimed at searching for UC~HII regions associated with extended emission. ", "introduction": "\\label{sec:1} Ultracompact (UC) HII regions are small (size $\\leq$ 0.1 pc), dense ($\\geq$ 10$^4$ cm$^{-3}$), photoionized hydrogen regions with high emission measure ($\\geq$ 10$^7$ {${\\rm pc\\ cm}^{-6}$}), surrounding recently formed ionizing OB type stars (e.g. Fig. 1a). These characteristics were observationally confirmed by \\cite{WC89} and \\cite{K94}, and more recent reviews are presented by \\cite{C02} and \\cite{LFR06}. The study of UC~HII regions began in 1967 via interferometric observations of compact HII regions (see \\cite{K02} for a summary). Because UC~HII regions are generally surrounded by a natal dust `cocoon', radio--continuum (RC) and infrared (IR) observations are needed to study them. In the RC, the first VLA surveys (e.g., \\cite{WC89}, \\cite{K94}) were made at 2 and 6 cm in configurations A and B, supplying arc--sec resolutions and sensitivities to structures up to 10--20$''$. The IR counterparts were mainly provided by IRAS, with resolutions of $\\sim$30$''$--2$'$. \\begin{figure} \\centering \\includegraphics[height=9.5cm]{edelafuente_fig1.eps} \\caption{The UC~HII region with EE G60.88--0.13 (IRAS 19442+2427). a) and b) VLA images in configuration B (the UC emission) and configuration C (the EE). c) Combined VLA Multi--Resolution--Clean (MRC) map. This map strongly suggests a direct connection between the UC and the extended emission. d) All IRAC bands image at meso--scales (gray) and superimposed contours from c). Dust is predominant in the region. A spatial location agreement between dust and the HII region is observed.} \\label{fig:1} % \\end{figure} Although the presence of large scale structures related to UC~HII regions (Fig. 1b and 8c of \\cite{K99}) had been inferred since 1967, the first interferometric surveys did not detect them. Nevertheless, their detection is possible with the VLA C and D configurations (e.g., \\cite{K99}), albeit at the expense of resolution towards the UC emission (UCE). To mitigate these spatial filtering effects, we made multi--configuration VLA observations to provide a multi-scale view of the UC~HII+EE (Fig.1c). Also, by using MSX observations, of higher resolution than IRAS, it is possible both to detect the EE and to resolve the UCE, since the infrared observations are sensitive to the full range of angular sizes. The best IR satellite observations available for this purpose (see Fig. 1d) are from the Spitzer Space Telescope ({\\it http://ssc.spitzer.caltech.edu}). ", "conclusions": "" }, "0710/0710.1773_arXiv.txt": { "abstract": "We investigate the influence of large-scale meridional circulation on solar p-modes by quasi-degenerate perturbation theory, as proposed by~\\inlinecite{lavely92}. As an input flow we use various models of stationary meridional circulation obeying the continuity equation. This flow perturbs the eigenmodes of an equilibrium model of the Sun. We derive the signatures of the meridional circulation in the frequency multiplets of solar p-modes. In most cases the meridional circulation leads to negative average frequency shifts of the multiplets. Further possible observable effects are briefly discussed. ", "introduction": "Meridional circulation is a large-scale flow observed on both hemispheres of the solar surface~\\cite{duvall79,hathaway96,komm93}. Its predominant direction is from the equator to the poles, and its amplitude is of the order of 15\\thinspace m/s. As mass does not accumulate in the polar regions a return flow from the poles to the equator is suspected deeper within the solar interior. The top half of the convection zone contains approximately 0.25\\% of the solar mass, the mass of the bottom half is approximately five times larger. Consequently, a poleward flow of 10\\thinspace m/s in the top half of the convection zone could be compensated by an equatorward flow of 2\\thinspace m/s in the lower half. The transport of magnetic flux from mid to low latitudes by such a flow at the bottom of the convection zone would last approximately 10\\thinspace years, which is close to the period of the solar magnetic cycle. In addition to magnetic flux, the meridional flow also transports angular momentum. Indeed, the circulation played a key role in early theories of the solar non-uniform rotation as well as of the magnetic cycle~\\cite{bjerknes26,kippenhahn63}. More recently, differential rotation has been explained in mean-field models as a consequence of the Reynolds stresses~\\cite{ruediger80,kueker01,kueker05}, and in three-dimensional numerical models by the influence of the Coriolis force on global convection~\\cite{miesch00,miesch06}. Nevertheless, meridional circulation occurs as well in these models, and in solar-cycle models it has regained popularity, since the mean-field latitude migration along the surfaces of isorotation that occurs in the traditional $\\alpha\\Omega$ dynamo~\\cite{parker55} does not seem to suffice. The effect of the circulation on the butterfly diagram had been demonstrated by ~\\inlinecite{roberts72}; it is considered to be essential in more recent versions of the $\\alpha\\Omega$ dynamo, which therefore have been termed `flux-transport dynamos'~\\cite{choudhuri95,dikpati99,nandy02,rempel06a,rempel06b}. Local helioseismology has investigated the strength of the meridional flow in the solar interior. By means of ring diagram analysis \\cite{hill88} \\inlinecite{haber02} inverted data for the circulation in a 15\\thinspace Mm deep region below the solar surface. In data from 1998 they found a flow emerging at high northern latitudes with equatorward orientation. This was interpreted as an evolving second cell of circulation. Further studies on the evolution of the flow either by ring-diagrams and by time-distance helioseismology (e.g.~\\opencite{zhao04},~\\opencite{zaatri06}) show predominantly a polward flow with a strong variability in the outer 15\\thinspace Mm of the Sun. The velocity reaches 40\\thinspace m/s. These findings were interpreted as the upper parts of meridional circulation cells in the two hemispheres. Theoretically, the influence of global-scale stationary flows on solar p-modes was studied in detail by quasi-degenerate perturbation theory~\\cite{lavely92}. These studies were successfully used to solve the forward problem for the influence of differential rotation on the p-modes. The results lead to an improved inversion method for determining the radial dependence of the differential rotation~\\cite{ritzwoller91}. Following~\\inlinecite{lavely92}, \\inlinecite{roth99} studied the influence of large-scale sectoral poloidal flow components that could be related with giant convection cells. They found that such flows yield additional frequency shifts that can only be described with quasi-degenerate perturbation theory, as these frequency shifts are effects of higher order. In a sequel study~\\inlinecite{roth02} were able to show that sectoral poloidal flows could only be found with the current inversion methods of global helioseismology as long as they exceed an amplitude of 10\\thinspace m/s. The meridional flow was found to be not detectable by the current inversion methods as the frequency splittings are fitted by an incomplete set of basis functions, which are tailored to measure only zonal toroidal flows. However, no detailed study on the effect of the meridional circulation on the oscillation frequencies was given, and few other attempts to derive observable signatures of the meridional flow in global helioseismology data exist~\\cite{woodard00}. In this contribution we concentrate on a theoretical study of the influence of the meridional circulation on the solar p-mode frequencies and describe a possible observable effect. This effect is significantly smaller than the frequency splitting caused by solar differential rotation. But as time series of global oscillation data exist that cover more than 10 years, the necessary frequency resolution might be available to detect it. The advantage of studying the meridional circulation by global helioseismic techniques is a possible inference of information from greater depths. ", "conclusions": "In this paper we used simple models of the meridional circulation to investigate their influences on the solar p-mode frequencies. The simplest model consisted of one cell per hemisphere and depth with a maximum horizontal flow velocity of 15\\thinspace m/s on the solar surface. The most complicated model we used had three cells per hemisphere and three revolution cells in depth with a horizontal flow velocity of 15\\thinspace m/s at the surface, too. We performed numerical calculations based on quasi-degenerate perturbation theory to obtain the frequency splittings of the solar p-modes due to this meridional circulation models. We find that the meridional circulation lifts the degeneracies of the multiplets. For the simplest model the shifts are on average only 0.01\\thinspace $\\mu$Hz with a few shifts up to several $\\mu$Hz. Models with more cells per hemisphere affect the p-modes stronger. For the model with three cells per hemisphere we find an average shift of 0.1\\thinspace $\\mu$Hz, with many shifts in the order of 1\\thinspace $\\mu$Hz. In most cases the shifts are negative due to the fact that the next neigboring mode in frequency dominates the shifting; and due to structure of the $l$-$\\nu$ diagram this nearest neighbor has usually a higher frequency causing therefore negative shifts. Comparing models with the same number of cells in latitude but different number of cells in depth we find only tiny differences in the resulting frequency splittings. In the case of $s=6$ the difference in the frequency splitting is on average 0.03\\thinspace $\\mu$Hz. However a number of differences in the order of 1\\thinspace $\\mu$Hz can occur. On the Sun the meridional circulation consists probably of a superposition of flow components with different numbers of cells in depth and latitude. In contrast to the models used in our work, the amplitudes of these flow are in reality likely to be very different and highly variable in time. Nevertheless, based on our results we are optimistic that the meridional circulation on the Sun could leave an observable signature in the p-mode frequencies. In order to detect the effect a frequency resolution of at least 0.1\\thinspace $\\mu$Hz must be achieved. This could be done by a several month long time series. In order to be able to distinguish not only the various components of the meridional circulation in the orders $s$ but also in the radial orders $n_c$, several years of data need to be averaged to obtain the required precision in the frequencies. As such long data sets exist it will be worth while to search for this effect in the p-mode frequencies. Compared to the splitting caused by the differential rotation the effect of meridional circulation is small. But as the effect of the rotational splitting is odd it can be separated from the even meridional splitting. However, e.g. asphericities and the magnetic field might cause symmetrical frequency shifts, too. Therefore, a lot of forward modelling will be necessary before the effect of the meridional circulation can be disentangled from these other effects. In this sense, one possible future extension of our work is the determination of the frequency shifts due to more sophisticated models of the meridional circulation, e.g. from three-dimensional numerical models. This might allow tailoring inversion routines for estimating the meridional circulation in the Sun from global solar oscillation frequencies." }, "0710/0710.3406_arXiv.txt": { "abstract": "{Observations of collimated outflows in young stellar objects indicate that several features of the jets can be understood by adopting the picture of a two-component outflow, wherein a central stellar component around the jet axis is surrounded by an extended disk-wind. The precise contribution of each component may depend on the intrinsic physical properties of the YSO-disk system as well as its evolutionary stage. }{ In this context, the present article starts a systematic investigation of two-component jet models via time-dependent simulations of two prototypical and complementary analytical solutions, each closely related to the properties of stellar-outflows and disk-winds. These models describe a meridionally and a radially self-similar exact solution of the steady-state, ideal hydromagnetic equations, respectively. }{ By using the PLUTO code to carry out the simulations, the study focuses on the topological stability of each of the two analytical solutions, which are successfully extended to all space by removing their singularities. In addition, their behavior and robustness over several physical and numerical modifications is extensively examined. Therefore, this work serves as the starting point for the analysis of the two-component jet simulations. }{ It is found that radially self-similar solutions (disk-winds) always reach a final steady-state while maintaining all their well-defined properties. The different ways to replace the singular part of the solution around the symmetry axis, being a first approximation towards a two-component outflow, lead to the appearance of a shock at the super-fast domain corresponding to the fast magnetosonic separatrix surface. These conclusions hold true independently of the numerical modifications and/or evolutionary constraints that the models have been undergone, such as starting with a sub-modified-fast initial solution or different types of heating/cooling assumptions. Furthermore, the final outcome of the simulations remains close enough to the initial analytical configurations showing thus, their topological stability. Conversely, the asymptotic configuration and the stability of meridionally self-similar models (stellar-winds) is related to the heating processes at the base of the wind. If the heating is modified by assuming a polytropic relation between density and pressure, a turbulent evolution is found. On the other hand, adiabatic conditions lead to the replacement of the outflow by an almost static atmosphere.}{} ", "introduction": "\\label{sec:intro} Observations made over the last two decades have shown that one class of the widespread astrophysical phenomenon of collimated plasma outflows (jets) is being launched from the vicinity of most young stellar objects (YSOs) (Burrows et al. \\cite{Bur96}). These supersonic mass outflows are found to be correlated with accretion (Cabrit et al. \\cite{Cab90}; Hartigan et al. \\cite{Har95}), to have narrow opening angles (Ray et al. \\cite{Ray96}) and to propagate for several orders of magnitude of spatial distances ranging from the AU to the pc scales (Dougados et al. \\cite{Dou00}; Hartigan et al. \\cite{Har04}). A central role in the launching, acceleration and collimation of these jets is widely believed to be played by magnetohydrodynamic (MHD) effects, which can also successfully remove the excessive angular momentum, allowing in this way the YSO to accrete and enter the main sequence. Nevertheless, although recent high angular resolution observations put several constraints on the different driving mechanisms proposed, it is not yet clear which is the dominant plasma launching mechanism in YSO jets. The system of a protostellar object basically contains two dynamical constituents, a central protostar and its surrounding accretion disk. Consequently, in Bogovalov \\& Tsinganos (\\cite{Bog01}) it is argued that jets observed from T Tauri stars most likely consist of two main steady components: (i) an inner pressure driven wind, which is non-collimated if the star is an inefficient magnetic rotator and (ii) an outer magneto-centrifugally driven disk-wind which provides most of the high mass loss rate observed. The relatively faster rotating magnetized disk produces the self-collimated wind which then forces all enclosed outflow from the central source to be collimated as well. This conclusion is confirmed by self-consistent simulations of the MHD equations. More recently, in Ferreira et al. (\\cite{Fer06}) it is argued that for the YSO jets observed in association with T Tauri stars, in addition to the pressure driven stellar outflow and the magneto-centrifugally launched extended ``warm'' disk-wind, a third component may be driven by magnetic processes at the magnetosphere/disk interaction, i.e., a sporadically ejected X-type wind. In addition, in Ferreira et al. (\\cite{Fer00}) a non steady ``two-flow'' scenario was also suggested, regarding a reconnection X- and a disk-wind. Nevertheless, the existence of such sporadic components is not supported by observational data as being one of the major contributors to the steady characteristics of jets, but rather could explain the observed variability in jet emission. On the other hand and within the same framework, recent observations (Edwards et al. \\cite{Edw06}; Kwan et al. \\cite{Kwa07}) show that both, disk and stellar winds are mainly present in T Tauri stars, with the dominant component being determined by the intrinsic physical properties of the particular YSO. From the analytical studies point of view, the complexity of the launching and collimation mechanisms of jets have forced researchers over the past several years to treat these two components separately. The only available analytical MHD models for jets are those characterized by the symmetries of radial and meridional self-similarity (Vlahakis \\& Tsinganos \\cite{Vla98}). In the former case, the solution is invariant as we look at a constant polar angle and in the latter, as we look at a constant spherical radius. The computational consequence of the respective symmetry is that by employing the separable spherical coordinates ($r$, $\\theta$), the set of coupled MHD equations reduces to a set of ordinary differential equations in $\\theta$, or, in $r$, respectively. The last remaining difficulty is to select solutions which are causally disconnected from the source of the outflow, i.e., those crossing the fast magnetosonic separatrix. In this way, one may construct either radially self-similar solutions closely related to the properties of magneto-centrifugally driven disk-winds (Blandford \\& Payne \\cite{Bla82}; Contopoulos \\& Lovelace \\cite{Con94}; Ferreira \\cite{Fer97}; Vlahakis et al. \\cite{Vla00}, hereafter VTST00), or meridionally self-similar ones to address pressure driven stellar outflows (Sauty \\& Tsinganos \\cite{Sau94}; Trussoni et al. \\cite{Tru97}; Sauty et al. \\cite{Sau02}, hereafter STT02). Since each self-similar symmetry corresponds to a particular component, we adopt the following initials: ADO (Analytical Disk Outflow) and ASO (Analytical Stellar Outflow) to refer to radially and meridionally self-similar solutions, respectively. Apart from the geometry, an intrinsic distinction between these two classes concerns the treatment of the energy equation. Nevertheless, the symmetry difference makes them complementary to each other, since the ADO solution becomes singular at the axis, whereas the ASO is by definition the proper one for modeling the area close to it. In addition, the properties of the launching region of the disk-wind, i.e., at large polar angles, are described more naturally by the ADO model. For a recent review on the analytical work on MHD outflows the reader is referred to Tsinganos (\\cite{Tsi07}). On the other hand, the increase of computational power along with the development of sophisticated numerical codes have allowed to study the time evolution of the MHD equations, giving us new perspectives of the physics involved. Jet launching and collimation have been mainly investigated with the following two methodologies: (i) by treating the disk as a boundary (Krasnopolsky et al. \\cite{Kra99}; Ouyed et al. \\cite{Ouy03}; Fendt \\cite{Fen06}) and (ii) by including the disk inside the computational box (Casse \\& Keppens \\cite{Cas04}; Meliani et al. \\cite{Mel06}; Zanni et al. \\cite{Zan07}). The former case allows a wider range of physical processes and mechanisms to be studied, whereas the latter, has the advantage of the jet evolution being consistent with that of the disk. However, with the exception of Gracia et al. (\\cite{Gra06}; hereafter GVT06), most numerical studies did not take advantage of the availability of the well studied analytical solutions which also allow a parametric study and therefore a better physical understanding of the problem of jet launching and collimation. The present work is the first attempt to numerically construct and study a two-component jet, by using as starting point the two well studied classes of analytical self-similar solutions. Towards this goal, in this paper we first address the question of the topological stability of each one of these two classes separately, before we combine them in the following paper. Concerning the ADO model, we shall use the VTST00 analytical solution, completing and considerably extending the GVT06 analysis. Therein, they found that the disk-wind model may attain a new steady-state configuration close enough to the initial analytical one, provided the assumption of some appropriate approximations around the axis. We further present the first numerical studies of ASO (meridionally self-similar winds), referring to the solutions of STT02, that are essential to model the region around the axis, just where the ADO model fails. Once the physical properties and stability of these two classes of solutions is clarified, our final aim will be to effectively build up a model that consistently merges the ASO and ADO solutions. Such simulations will be presented in a future work where we will study the launching and propagation of a collimated stellar wind around the system axis surrounded by a disk wind. Finally, a word on the term topological stability used in this paper. Classical stability theory addresses the question whether a given equilibrium configuration evolves away from (=unstable) or back to (=stable) the initial equilibrium when perturbed. In the present context, topological (or structural) stability refers to the question whether a given configuration preserves its topological properties when subject to various perturbations. Needless to say, that topologically unstable configurations may well be stable from the classical point of view and vice versa. The paper is structured as follows: In section \\S\\ref{sec:anmod} the formalism of the ADO and ASO solutions is briefly presented. In section \\S\\ref{numst} their implementation is explained and the numerical models to be investigated are presented. Section \\S\\ref{sec:res} reports the results obtained by carrying out the respective simulations. Finally in section \\S\\ref{sec:disc} we discuss our results in the framework of the future matching and we report the conclusions of this work. ", "conclusions": "\\label{sec:disc} In this paper we have studied several physical and numerical aspects concerning two classes of the self-similar models, each associated with a disk- and a stellar-wind, in the framework of the upcoming work that combines them to describe a two-component outflow. These analytical solutions (ADO and ASO) were appropriately modified, implemented as initial conditions and evolved in time. Our main conclusions are the following: \\begin{itemize} \\item The {\\it Analytical Disk Outflow (radially self-similar) solution} has been successfully validated for its stability and robustness against several physical and numerical issues. This argument holds true even though the analytical solution was in many cases significantly modified. We have constructed numerical models and carried out simulations a) by assuming the extreme cases of an isothermal and an adiabatic evolution, b) by treating the diverging behavior of the solution at the axis with different kinds of extrapolation schemes, mimicking a stellar wind component and c) by changing the size, resolution and geometry of the computational box. In all cases, the poloidal critical surfaces, with the exception of the FMSS, were not readjusted, but rather matched perfectly to their initial position. The numerical solution always maintained the property of the successful crossing of all three critical surfaces producing an outflow causally disconnected from the base. This is achieved with the formation of a shock, corresponding to the numerically readjusted FMSS. In particular, this shock acts as a ``wall'' protecting the sub-modified-fast magnetosonic regions (source regions of the disk wind) from any perturbations taking place due to the modification of the models close to the axis (e.g. an effective stellar wind). However, the numerically readjusted FMSS (shock) does not coincide with the analytical one, with this departure being dictated by the respective numerical modifications of the models under consideration. A highly significant result is the fact that such a conclusion holds true even if we initialize the simulation with a sub-modified-fast solution, i.e. a solution with its whole domain causally connected. We found that, during the simulation, such a numerical model self-adapts to produce a shock (corresponding to the FMSS), hence no information of the downstream region can travel back to affect the launching region. This implies that even MHD outflow solutions, that do not successfully cross all three critical points, will probably converge to ``astrophysically correct'' solutions once evolved in time (see also Ferreira \\cite{Fer97}). On the other hand, the study of GVT06 was successfully extended down to the equator with the help of simulations using spherical coordinates. Furthermore, by adopting different assumptions for the energy source terms, it was shown that the solution is only slightly and accordingly self-modified maintaining all its well defined properties. This is in agreement with the fact that the ADO solution describes essentially a magneto-centrifugally accelerated outflow. \\item The {\\it Analytical Stellar Outflow (meridionally self-similar) solution}, which was validated in time-dependent simulations for the first time, maintained its well-defined equilibrium as expected. Such conclusion is supported by simulations performed with both the super- and sub-Alfv\\'enic regions included. Quite critical are, contrary to disk winds, the effects of the energetics in such thermally driven models. Although different assumptions of the energy equation in the super-Alfv\\'enic domain did not yield any significant modification of the analytical solution, strong variations of the structure of the axial outflows are found if modifications of the heating/cooling mechanisms occur in the initial accelerating region. In particular a polytropic assumption, mimicking isothermal conditions, would produce a turbulent weak outflow, while an adiabatic evolution asymptotically reaches a static atmosphere. We are tempted to relate the heating intermittency and even a switching off in such an ASO solution with the observed variability of accretion-driven YSO outflows. \\item All previous statements hold true while being in perfect agreement with physically consistent requirements, such as specifying the correct type of boundary conditions: a) according to the propagation direction of the MHD waves, b) the axisymmetry holding around the axis and c) the constancy of certain physical variables at both a conical surface close to the equatorial plane and a radial one close to the origin, implying the presence of an underlying disk and the stellar surface, respectively. Therefore, the results can safely be trusted since they are not subject to any artificial forcing. \\item Last, but certainly not least, is the fact that almost all models of both classes reached a steady- or quasi-steady-state. In this context, the upcoming mixing of the two complementary classes of self-similar solutions in order to study a two-component jet is well founded and promising. The final numerical model will incorporate both proposed scenarios of a pressure-driven outflow (ASO) surrounded by an extended, magneto-centrifugally driven disk wind (ADO). This task is undertaken in the second paper of this series. \\end{itemize}" }, "0710/0710.1918_arXiv.txt": { "abstract": "We examine production of Li on the surface of a low-mass secondary in a black hole soft X-ray transient (BHSXT) through the spallation of CNO nuclei by neutrons which are ejected from a hot (> 10 MeV) advection-dominated accretion flow (ADAF) around the black hole. Using updated binary parameters, cross sections of neutron-induced spallation reactions, and mass accretion rates in ADAF derived from the spectrum fitting of multi-wavelength observations of quiescent BHSXTs, we obtain the equilibrium abundances of Li by equating the production rate of Li and the mass transfer rate through accretion to the black hole. The resulting abundances are found to be in good agreement with the observed values in seven BHSXTs. We note that the abundances vary in a timescale longer than a few months in our model. Moreover, the isotopic ratio \\nuc{Li}{6}/\\nuc{Li}{7} is calculated to be about 0.7--0.8 on the secondaries, which is much higher than the ratio measured in meteorites. Detection of such a high value is favorable to the production of Li via spallation and the existence of a hot accretion flow, rather than an accretion disk corona system in quiescent BHSXT. ", "introduction": "High abundances of Li have been detected in late-type secondaries of black hole soft X-ray transients (BHSXTs) and a neutron star soft X-ray transient (NSSXT) in quiescence~\\citep{martin92,martin94,martin96}, though Li would be destructed in a deep convective envelope of a late-type star. The Li enrichment has not, however, been observed on a late-type secondary in a compact binary with a white dwarf~\\citep{martin95}. These facts strongly suggest that a production mechanism of Li operates in compact binaries~\\citep{yn97,gk99} and that the nature of the primaries is crucial for the mechanism, though rotation might reduce the destruction of Li in the envelope of the secondary~\\citep{mjs05}. Multi-wavelength spectra of BHSXTs in quiescence are successfully fitted to the radiation from an advection-dominated accretion flow (ADAF) around the black hole~\\citep{nmy96,nbm97}. Density is so low in ADAF, that ions interact inefficiently with electrons. Consequently ions have high temperatures due to viscous heating up to about 30 MeV near the inner edge of ADAF. At such high temperatures, $\\alpha$-$\\alpha$ reaction proceeds to synthesize Li inside ADAF~\\citep{martin94,yn97}. It is necessary that a fraction $10^{-3}-10^{-4}$ of the accreting gas is transported to the secondary to explain the high abundances of Li observed in BHSXTs. However, such a high fraction is uncertain to be realized due to strong gravity of the black hole and the Coulomb interactions with nuclei inside ADAF~\\citep{gk99}. Helium breaks via spallation with protons to produce neutrons at the inner region of ADAF. A large fraction of neutrons can be ejected from ADAF, because they do not interact with nuclei through the Coulomb interactions. Neutrons intercepted by the secondary interact with CNO nuclei through spallation to produce Li on the surface~\\citep{gk99}. This scenario is of particular interest, because the Li enrichment is anticipated in secondaries only for BHSXTs and NSSXTs, but for white dwarfs as primaries where ADAF cannot attain enough high temperatures to break helium to nucleons there. In the present paper, we evaluate the Li abundances on the surface of secondaries in BHSXTs, following the scenario proposed by \\citet{gk99}. To this end, we use updated binary parameters, such as the mass $M$ of a black hole, the mass $M_*$ and radius $R_*$ of a secondary, mass accretion rates derived from the spectrum fitting of multi-wavelength observation of BHSXTs in quiescence, and cross sections of neutron-induced spallation reactions. Then, we compare the resulting abundances with the observed values, and show that the agreement is quite well. Moreover, we predict the isotopic ratio \\nuc{Li}{6}/\\nuc{Li}{7} on the secondaries in BHSXTs. ", "conclusions": "" }, "0710/0710.4226_arXiv.txt": { "abstract": "We report on simultaneous optical and X-ray observations of the Seyfert galaxy, NGC~3147. The XMM-\\textit{Newton} spectrum shows that the source is unabsorbed in the X-rays ($N_H<5\\times10^{20}$ cm$^{-2}$). On the other hand, no broad lines are present in the optical spectrum. The origin of this optical/X-rays misclassification (with respect to the Unification Model) cannot be attributed to variability, since the observations in the two bands are simultaneous. Moreover, a Compton-thick nature of the object can be rejected on the basis of the low equivalent width of the iron K$\\alpha$ line ($\\simeq130$ eV) and the large ratio between the 2-10 keV and the [O\\textsc{iii}] fluxes. It seems therefore inescapable to conclude that NGC~3147 intrinsically lacks the Broad Line Region (BLR), making it the first ``true'' Seyfert 2. ", "introduction": "The basic assumption of Unified Models of Active Galactic Nuclei (AGN) is that type 1 and type 2 objects are intrinsically the same, the apparent difference being solely due to orientation effects \\citep[e.g.][]{antonucci93}. The absorbing medium assumes the fundamental role in this scenario. It is usually envisaged as an optically thick `torus', embedding the nucleus and the Broad Line Region (BLR). If we observe the torus edge-on, all the nuclear radiation, as well as the broad optical lines coming from the BLR, is completely blocked and we classify the source as a type 2. The narrow lines are still visible, because the Narrow Line Region (NLR) is located farther away from the nucleus, beyond the torus. On the other hand, if the torus does not intercept our line of sight, we observe every component of the spectrum and the object is classified as a type 1. Simple extrapolations of the optical/UV scenario to the X-ray emission do not, however, always easily fit the observations, leading sometimes to different classifications between the two bands. A larger amount of absorbing material is usually measured from X-ray observations in Seyfert 2s with respect to Seyfert 1s, as expected \\citep[e.g.][]{awaki91,ris99}. However, a number of ``true'' Seyfert 2 candidates have been found, i.e. objects with no absorption in the X-rays and no broad optical lines \\citep[e.g.][]{pappa01,pb02,wol05}, or with a large intrinsic Balmer decrement of the BLR \\citep{bcc03,corr05}. The so-called `naked' AGN may be similar objects, being characterised by the absence of broad lines, but strong variability in the optical band \\citep{hawk04} and, apparently, no absorption in the X-rays \\citep{gliozzi07}. These findings represent a challenge to the Unified Models and may require new classes of objects with intrinsic differences, like the absence of BLR. Nevertheless, these sources may be highly variable and may change their optical and/or X-ray appearance in different observations. These `changing-look' AGN are not uncommon. In some cases, this behaviour is best explained by a real `switching-off' of the nucleus \\citep[see e.g.][]{mgm03,gua05}, in others by a variable column density of the absorber \\citep[e.g.][]{elvis04,ris05}. If the optical and the X-ray spectrum are taken in two different states of the source, it is clear that the disagreement between the two classifications may be only apparent. Therefore, the key to find genuine `unabsorbed Seyfert 2s' is represented by simultaneous X-ray and optical observations. NGC~3147 (z=0.00941) belongs to the Palomar optical spectroscopic survey of nearby galaxies \\citep{hfs95}. Its classification as a Seyfert 2 is based on both the relative strength of the low-ionization optical forbidden lines with respect to the hydrogen Balmer lines and the ratio of [O\\textsc{iii}] to H$\\beta$, together with the lack of broad permitted lines \\citep{hfs97}. \\textit{ASCA} provided the first X-ray spectrum, which appeared Seyfert 1-like, without significant absorption and a standard powerlaw index \\citep{ptak96}. The lack of obscuration was later confirmed by \\textit{BeppoSAX} \\citep{dad07} and \\textit{Chandra}, which also showed that no off-nuclear source can significantly contribute to the nuclear emission \\citep{tw03}. In order to reconcile the X-ray data with the optical classification, \\citet{ptak96} suggested that NGC~3147 was a Compton-thick source, given also the relatively large equivalent width (EW) of the iron line. However, this hypothesis was rejected by the use of diagnostic diagrams based on $F_\\mathrm{X}/F_{\\mathrm{[OIII]}}$ and $F_\\mathrm{X}/F_{\\mathrm{IR}}$ ratios \\citep{pb02}. NGC~3147 is, therefore, a genuine candidate to be an unabsorbed Seyfert 2 galaxy. In this paper, we report on simultaneous XMM-\\textit{Newton} and optical observations of NCG~3147, in order to settle the issue. In the following, errors correspond to the 90\\% confidence level for one interesting parameter ($\\Delta \\chi^2 =2.71$), where not otherwise stated. The adopted cosmological parameters are $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_\\Lambda=0.73$ and $\\Omega_m=0.27$ (i.e. the default ones in \\textsc {Xspec} 12.3.1). ", "conclusions": "The best fit model for the XMM-\\textit{Newton} spectra is that of a typical type 1 AGN. The amount of neutral column density measured in excess of the Galactic one is of the same order of the latter, and it is therefore fully consistent with being due to the host galaxy's interstellar medium. The powerlaw index, although somewhat flatter than the average type 1 radio-quiet object, is well within the large dispersion of the hard X-ray $\\Gamma$ distribution \\citep[e.g.][]{pico05}. The EW of the neutral iron K$\\alpha$ line is fully consistent with those typically found in unobscured AGN, especially given its low luminosity \\citep[e.g.][]{bianchi07}. The only peculiar aspect of its X-ray spectrum is represented by the Fe \\textsc{xxv} emission line, whose EW is larger than the one typically arising in a Compton-thin, photoionised gas \\citep[see e.g.][]{bm02,bianchi05}. However, the line is not statistically very significant and the errors on the EW are quite large (see Sect. \\ref{xmmanalysis}). To test further the consistency of NGC 3147 being a type 1 AGN, we used the QSO template from \\citet{elvis94} and scaled it to the X-ray flux measured by our X-ray observation. The predicted continuum level at optical wavelengths is fully consistent with the measured nuclear spectrum. In addition, we have examined the Optical Monitor \\citep{Mason01} data of our XMM-\\textit{Newton} observation, where the nucleus of NGC~3147 is detected as a point source. Once deconvolved the disk contribution and corrected for Galactic reddening, we get fluxes of $(1.81\\pm 0.14)$ and $(0.94\\pm0.14)\\times 10^{-15}$ erg cm$^{-2}$ s$^{-1}$ \\AA, at 2910 and 2310 \\AA, respectively, leading to a two-point spectral index $\\alpha_{ox}\\simeq-1.33$\\footnote{We adopt the original definition by \\citet{tan79} as the ratio between X-ray and UV flux densities: $\\alpha_{ox}=-0.384 \\log[f(2\\,\\mathrm{keV})/f(2500\\,\\mathrm{\\AA})]$.}, as found for most Seyfert 1s \\citep[e.g.][]{stra05}. However, the analysis of the \\textit{simultaneous} optical spectrum presented in this paper confirms the lack of broad permitted lines in this source. The very low column density measured in the X-rays ($N_H<5\\times10^{20}$ cm$^{-2}$ at the 99\\% confidence level for two interesting parameters, corresponding to $\\mathrm{A_V}<0.3$) is at odds with the large amount of dust required to obscure the BLR. Generally, it is found that any deviation from the Galactic gas-to-dust ratio in AGN always goes in the opposite direction: the obscuration by dust is \\textit{lower} than what expected from the associated gas column density \\citep{mai01}. It is difficult to find a physical mechanism able to suppress gas without destroying dust. If the source were Compton-thick, the lack of X-ray absorption could be apparent, because we would not observe the primary emission at all, but only the reflected light, unaffected by any line-of-sight absorption. This solution is untenable for NGC~3147, because of the EW of the neutral iron line, much smaller than the one expected in the case of a reflection-dominated object \\citep[$\\simeq1$ keV: see e.g.][]{mbf96}. However, the (relatively) large EW of the Fe\\textsc{xxv} line may be a signature of reflection from ionised material. In this case, the reprocessed spectrum would be indistinguishable from the primary powerlaw, but an Fe\\textsc{xxv} EW as low as some hundreds of eV would be likely accompanied by detectable emission from Fe\\textsc{xxvi} or lower ionised species \\citep[see e.g.][]{bm02}. This solution would require a rather ad hoc geometry: a Compton-thick neutral absorber blocks the primary emission, while the reprocessed spectrum is instead dominated by an highly ionised material. Moreover, the high ratio between the 2-10 keV and the [O\\textsc{iii}] reddening-corrected flux ($\\simeq20$) strongly rejects the hypothesis that we are not directly observing the primary emission \\citep[in this case, a ratio lower than 1 is expected: see e.g.][and references therein]{pb02}. Therefore, the source is quite unlikely to be Compton-thick. The only solution left is that the source intrinsically lacks the BLR. In this respect, it is interesting to note that NGC~3147 has no hidden BLR (HBLR) in polarised light (Tran, private communication). It was shown that HBLR sources are characterized, on average, by larger luminosities than non-HBLR ones \\citep[e.g.][]{gh02}. Recently, \\citet{es06} presented a model which depicts the torus as the inner region of a clumpy wind outflowing from the accretion disc. A key prediction of this scenario is the disappearance of the BLR at very low bolometric luminosities. Adopting a constant bolometric correction of 20 \\citep[][but any other choice is irrelevant for the present estimate]{elvis94} to the observed X-ray luminosity, the bolometric luminosity of NGC~3147 is about $5\\times10^{42}$ erg s$^{-1}$, which is somewhat larger than the `threshold' calculated by \\citet{es06} and, in any case, larger than other low-luminosity objects showing broad lines \\citep{ho97}. Moreover, the disappearance of the BLR \\textit{follows} that of the torus, at lower luminosities. Therefore, their model predicts the absence of the torus in NGC~3147, requiring the strong neutral iron line to be produced in the accretion disc. Current data cannot confirm nor reject this hypothesis. Another possibility is that the BLR is intimately linked to the accretion rate, rather than the luminosity, of the AGN, being formed in accretion disk instabilities occurring around a critical radius \\citep{nic00}. Below a minimum accretion rate, this radius falls below the innermost stable orbit and the BLR cannot form. We found in the recent literature two estimates for the black hole mass of NGC~3147: $6.2\\times10^{8}$ M$_{\\odot}$ \\citep[][inferred from the mass-velocity dispersion correlation]{merl03} and $2.0\\times10^{8}$ M$_{\\odot}$ \\citep[][using a mass-$K_s$ bulge luminosity relation]{dd06}. Combined with the above-derived bolometric luminosity, we get a very low Eddington rate for NGC~3147, roughly ranging between $8\\times10^{-5}$ and $2\\times10^{-4}$. In any case, this is well below the threshold (around $1\\times10^{-3}$) proposed by \\citet{nic00} and \\citet{nic03}, thus supporting this scenario. In conclusion, it seems inescapable that the key parameter for the observational properties of this source is not orientation, but an intrinsic feature (be it low accretion rate or luminosity), which prevents the BLR to form. NGC~3147 is therefore a ``true'' Seyfert 2 without the BLR, the first unambiguous component of a new class of AGN which requires to be fitted in a more general Unified Model." }, "0710/0710.0319_arXiv.txt": { "abstract": "We present new evolutionary models for Type Ia supernova (SN Ia) progenitors, introducing mass-stripping effect on a main-sequence (MS) or slightly evolved companion star by winds from a mass-accreting white dwarf (WD). The mass-stripping attenuates the rate of mass transfer from the companion to the WD. As a result, quite a massive MS companion can avoid forming a common envelope and increase the WD mass up to the SN Ia explosion. Including the mass-stripping effect, we follow binary evolutions of various WD + MS systems and obtain the parameter region in the initial donor mass -- orbital period plane where SNe Ia occur. The newly obtained SN Ia region extends to donor masses of $6-7 ~M_\\sun$, although its extension depends on the efficiency of mass-stripping effect. The stripped matter would mainly be distributed on the orbital plane and form very massive circumstellar matter (CSM) around the SN Ia progenitor. It can explain massive CSM around SNe Ia/IIn(IIa) 2002ic and 2005gj as well as tenuous CSM around normal SN Ia 2006X. Our new model suggests the presence of very young ($\\lesssim 10^8$~yr) populations of SNe Ia, being consistent with recent observational indications of young population SNe Ia. ", "introduction": "The nature of Type Ia supernova (SN Ia) progenitors has not been clarified yet \\citep[e.g.,][]{nie04, nom00}, although it has been commonly agreed that the exploding star is a mass-accreting carbon-oxygen white dwarf (C+O WD). For the exploding WD itself, the observed features of SNe Ia are better explained by the Chandrasekhar mass model than the sub-Chandrasekhar mass model \\citep[e.g.,][]{liv00}. However, there has been no clear observational indication as to how the WD mass gets close enough to the Chandrasekhar mass for carbon ignition; i.e., whether the WD accretes H/He-rich matter from its binary companion [single degenerate (SD) scenario], or two C+O WDs merge [double degenerate (DD) scenario]. Recently, the following two important findings have been reported in relation to the SN Ia progenitors: (1) circumstellar matter (CSM) around the progenitors, and (2) a very young ($\\lesssim 10^8$~yr) population of the progenitors. {\\bf Circumstellar Matter:} In the SD scenario, H/He-rich CSM is expected to exist around SNe Ia as a result of mass transfer from the companion as well as the WD winds \\citep*[e.g.,][]{nom82, hkn99}. Thus searching for H/He-rich CSM is one of the key observations to identify the progenitors \\citep[e.g.,][]{lun03}. Recently detections of such CSM have been reported for several SNe Ia, i.e., observations of narrow H-emission lines in SNe 2002ic \\citep{hau03} and 2005gj \\citep{ald06,pri07} (Type Ia/IIn or IIa \\citep{den04}), thermal X-rays from 2005ke \\citep{imm06}, and \\ion{Na}{1}~D lines in 2006X \\citep{pat07a}. The identification of SN 2002ic as an SN Ia has been confirmed by the recent spectral comparison between SN 2005gj and SNe Ia \\citep{pri07}, being against the Type Ic suggestion by \\citet{ben06}. Several CSM interaction models suggested a $1 - 2 ~M_\\sun$ CSM \\citep{chu04,nom05}. The evolutionary origin of such a massive CSM has been explored by \\citet{liv03} based on a common envelope evolution model, by \\citet{han06} from the delayed dynamical instability model of binary mass transfer, and by \\citet{woo06} based on a recurrent nova model with a red giant companion. For normal SNe Ia, non-detection of radio has put the upper limit of mass loss rate as $\\dot M / v_{10} \\lesssim 10^{-8} M_\\sun$~yr$^{-1}$, where $v_{10} \\equiv v / 10$~km~s$^{-1}$ \\citep{pan06}. However, the optical observations of SN 2006X have detected variable Na I D lines from CSM, whose expansion velocity and mass have been estimated to be $v_{10} \\sim$ 10 and $\\sim 10^{-4} ~M_\\sun$ \\citep{pat07a}. Patat et al. have suggested that the CSM in SN 2006X originated from the red-giant companion because of relatively low velocities. Comparing the SN 2006X light curves with the other normal SNe Ia light curves, \\citet{wan07a} suggested that the obvious deviation, the decline rate is slowing down in a later phase, can be explained by an interaction between ejecta and CSM or a light echo of circumstellar/interstellar matter \\citep[see also][]{wan07b}. {\\bf Young Population:} According to \\citet{man06}, the present observational data of SNe Ia are best matched by a bimodal population of the progenitors, in which about 50 percent of SNe Ia explode soon after their stellar birth at the {\\sl delay time} of $t_{\\rm delay} \\sim 10^8$ yr, while the remaining 50 percent have a much wider distribution of the {\\sl delay time} of $t_{\\rm delay} \\sim$ 3 Gyr. \\citet{aub07} recently reported evidence for a short (less than 70 Myr) delay time component in the SN Ia population. In this paper, we define the term {\\sl delay time} as {\\sl the age of a binary system at the SN Ia explosion}, in order to compare our results with the earlier results \\citep[e.g.,][]{gre83, gre05, man06}. \\begin{figure*} \\epsscale{0.6} \\plotone{f1.eps} \\caption{ A schematic configuration of a binary evolution including mass-stripping effect. (a) Here we start a pair of a C+O WD and a more massive main-sequence (MS) star with a separation of several to a few tens of solar radii. (b) When the secondary evolves to fill its Roche lobe, mass transfer onto the WD begins. The mass transfer rate exceeds a critical rate for optically thick winds. Strong winds blow from the WD. (c) The hot wind from the WD hits the secondary and strips off its surface. (d) Such stripped-off material forms a massive circumstellar disk or torus and it gradually expands with an outward velocity of $\\sim 10-100$~km~s$^{-1}$. The interaction between the WD wind and the circumstellar torus forms an hourglass structure. The WD mass increases up to $M_{\\rm Ia}= 1.38 ~M_\\sun$ and explodes as an SN Ia. When we observe the SN Ia from a high inclination angle such as denoted by ``line of sight,'' circumstellar matter (CSM) can be detected as absorption lines like in SN 2006X. \\label{stripping_evolution}} \\end{figure*} This kind of short delay times ($t_{\\rm delay} \\lesssim 10^8$~yr) of SNe Ia have been suggested from the distribution of SNe Ia relative to spiral arms \\citep[e.g.,][]{bar94, del94}. Recently, \\citet{dis03} reported, based on the {\\it Chandra} data from four external galaxies: an elliptical galaxy (NGC 4967), two face-on spiral galaxies (M101 and M83), and an interacting galaxy (M51), that in every galaxy there are at least several hundred luminous supersoft X-ray sources (SSXSs) with a luminosity of $\\gtrsim 10^{37}$ erg~s$^{-1}$ and that, in the spiral galaxies M101, M83, and M51, SSXSs appear to be associated with the spiral arms. The latter may indicate that SSXSs are young systems, possibly younger than $10^8$ yr, and has some close relation to the young population of SNe Ia. The SD scenario has ever not predicted such young populations of $t_{\\rm delay} \\sim 10^8$ yr, corresponding to, at least, the zero-age main-sequence (ZAMS) stars at mass $5-6 ~M_\\sun$ \\citep[see, e.g.,][]{lih97, hknu99, lan00, han04}. In the present paper, we propose a scenario for such a young SN Ia population by introducing {\\sl mass-stripping effect} into binary evolutions. Mass-accreting WDs blow optically thick winds when the mass transfer rate to the WD exceeds the critical rate of ${\\dot M}_{\\rm cr} \\sim 1 \\times 10^{-6} M_\\sun$~yr$^{-1}$ \\citep{hkn96}. The WD wind collides with the secondary's surface and strips off matter. When the mass-stripping effect is efficient enough, the mass transfer rate to the WD is attenuated and the binary can avoid the formation of a common envelope even for a rather massive secondary. The mass-stripping effect on a MS companion has been first introduced by \\citet{hac03kb, hac03kc}, who analyzed two quasi-periodic transient supersoft X-ray sources, RX~J0513.9$-6951$ and V~Sge: RX~J0513 shows a quasi-periodic oscillation between optical high ($\\sim 100-120$ days) and low ($\\sim 40$ days) states with an amplitude of 1 mag \\citep{alc96}. RX~J0513 is X-ray bright only during the optical low states \\citep{rei00}. \\citet{hac03kb} proposed a model that the mass transfer is modulated by the WD wind because the WD wind collides with the companion and strips off its surface and attenuates the mass transfer rate. When the mass transfer rate decreases below the critical rate ${\\dot M}_{\\rm cr}$, the WD wind stops and supersoft X-ray turns on. This corresponds to an optical low sate. Then the mass-transfer rate recovers because of no attenuation by WD winds and the WD blows winds again. X-ray turns off and an optical high state resumes and the binary starts the next cycle of quasi-periodic oscillation. Such a self-sustained model naturally explains major characteristics of quasi-periodic high and low states and this success encourages us to adopt the same idea in the evolution scenario of supersoft X-ray sources and SN Ia progenitors. In the present paper, we show that this mass-stripping effect derives (1) formations of circumstellar matter (CSM) around SNe Ia and (2) a very young population of SNe Ia. We summarize our basic treatments of mass-stripping effect and binary evolutions in \\S 2, and then show our numerical results and their relations to a very young population of SNe Ia in \\S 3. In \\S 4 we present the origin of CSM around SNe Ia based on our results and show a relation between the very young population of SNe Ia and their massive CSM. Discussion and concluding remarks follow in \\S\\S 5 and 6. \\begin{figure*} \\epsscale{0.7} \\plotone{f2.epsi} \\caption{ SN Ia evolutions for two typical cases of WIND and CALM. (a) Case WIND: starting from $M_{\\rm WD,0} =1.0 ~M_\\sun$, $M_{2,0} =5.0 ~M_\\sun$, and $P_0 = 2.15$ days with $c_1=3$, the WD reaches the SN Ia explosion in the wind phase at $t=6.57 \\times 10^5$ yr. The WD mass ($M_{\\rm WD}$), secondary mass ($M_2$), mass loss rate from the secondary (${\\dot M}_2$), WD wind mass loss rate (${\\dot M}_{\\rm wind}$), radius of the secondary ($R_2$), effective radius of the Roche lobe for the secondary ($R_2^*$), and orbital period ($P_{\\rm orb}$) are plotted. Only the orbital period is multiplied by four to easily see its change. (b) Case CALM: starting from $M_{\\rm WD,0} =1.0 ~M_\\sun$, $M_{2,0} =5.0 ~M_\\sun$, and $P_0 = 6.79$ days with $c_1=3$, the WD reaches the SN Ia explosion but in an SSXS phase without winds at $t= 6.93 \\times 10^5$ yr. The WD wind stops at $t=5.5 \\times 10^5$~yr. Even after that, the WD loses its mass due to weak helium shell flashes \\citep{kat99h}. Here ${\\dot M}_{\\rm wind}$ includes an average mass loss rate by helium shell flashes and thus does not become zero after the optically thick wind of steady hydrogen shell burning stops. Values of the secondary radius ($R_2$) and the Roche lobe radius for the secondary ($R_2^*$) are divided by two to squeeze them into the figure. \\label{evolution_sn2005gj}} \\end{figure*} ", "conclusions": "Both Cases WIND and CALM originate from the systems with massive donors, i.e., young population. It would be important to make some comparisons with the observational data, such as frequency and population. The red hatched regions in Figures \\ref{zams_evl_con_c3_m110}, \\ref{zams_evl_con_c3_m100}, and \\ref{zams_evl_con_c3_m090} indicate a region in which the progenitor explodes at $t_{\\rm delay} \\le 100$~Myr. Also the dashed line and the dotted lines correspond to $t_{\\rm delay} =$ 200~Myr and 400~Myr, respectively. We see in Figure \\ref{birth_rate_population_ms} that Case WIND and thus SNe Ia/IIn (IIa) are realized by the very young system with $t_{\\rm delay} \\lesssim 100-200$~Myr. If $M_{\\rm WD, 0} \\lesssim 0.9 ~M_\\sun$, we have almost no region of Case WIND, different from the cases of $M_{\\rm WD, 0} \\gtrsim 1.0 ~M_\\sun$. If all the WD + MS system with $M_{2,0} \\gtrsim 3-6 ~M_\\sun$ ($c_1=3$), $M_{\\rm WD,0} \\gtrsim 1.0 ~M_\\sun$ ($M_{1,0} \\gtrsim 6.5~M_\\sun$), and $P_0 \\sim 0.5 - 2$~days produces SNe Ia/IIn (IIa) events (Table \\ref{condition_sn1a_explosion}), the frequency of these events is estimated to be $\\sim 5$\\% (including both the WD + MS and WD + RG systems with their total number ratio of 4:2). A group of Type IIn SNe such as SNe 1997cy and 1999E show a very similar spectroscopic and photometric features to SN 2002ic \\citep{wan04, den04, pri07}. If these are in fact all Type Ia/IIn (IIa) SNe, their frequency can be estimated to be $\\sim 5_{-4}^{+7}$~\\% \\citep{pri07}, which is consistent with the above estimate. Type Ia supernovae play a key role in astrophysics, and thus our progenitor model has important implications. Our model depends essentially on the parameter of stripping effect, $c_1$, which depends on the properties of WD winds, such as asphericity, velocities, and the efficiency of energy conversion. Also we calculate the mass transfer rate using the simple approximate binary models. In order to improve these parameterization and approximations, we need multi-dimensional hydrodynamical simulations, which are beyond the scope of the present study. In the present approach, we constrain the $c_1$ parameter observationally, and estimate $c_1 \\sim 7-8$ and $c_1 \\sim 1.5-10$ from the analysis of V~Sge and RX~J$0513.9-6951$, respectively. With keeping in mind the necessity of further theoretical and observational studies to confirm our new progenitor systems, we summarize the basic results of our new SN Ia scenario: (1) Mass-accreting WDs blow an optically thick wind when the mass transfer rate to the WD exceeds the critical rate of ${\\dot M}_{\\rm cr} \\sim 1 \\times 10^{-6} M_\\sun$~yr$^{-1}$. The WD wind collides with the secondary's surface and strips off its surface. If the mass-stripping effect is efficient enough, the mass transfer rate to the WD is attenuated and the binary can avoid formation of a common envelope even for a rather massive secondary. Including this mass-stripping effect into our binary evolution model of the WD + MS systems, we have found a new evolutionary scenario, in which a companion as massive as 6--$7~M_\\sun$ can produce an SN Ia for a reasonable strength of mass-stripping effect, say $c_1 \\sim 3$. (2) We have followed simplified binary evolutions and obtained the SN Ia region in the $\\log P_0$--$M_{2,0}$ (initial orbital period -- initial donor mass) plane. The newly obtained SN Ia region extends to massive donor masses up to $M_{2,0} \\sim 6-7 ~M_\\sun$ for $P_0 \\sim 0.5-10$ days, although its extension depends on the strength of mass-stripping effect, $c_1$, i.e., $M_{2,0} \\sim 7-8 ~M_\\sun$ for $c_1=10$, $M_{2,0} \\sim 5-6 ~M_\\sun$ for $c_1=3$, and $M_{2,0} \\sim 4 ~M_\\sun$ for $c_1=1$. (3) We have estimated that the SN Ia birth rate in our Galaxy is $\\nu_{\\rm WD+MS} \\sim 0.004$~yr$^{-1}$ (for $c_1 = 3$), which is consistent with the observation. The rates of young populations, i.e., $t_{\\rm delay} \\le 100$~Myr and $t_{\\rm delay} \\le 200$~Myr, are about 50\\% and 80\\% of the total SN Ia rate of the WD + MS channel. These short delay times of SN Ia progenitors are consistent with the recent observational suggestions that a half of SNe Ia belong to such a very young population as the delay time of $t_{\\rm delay} \\sim 10^8$~yr. (4) Another channel of the WD + RG system shows a broad distribution of the delay time over 2--3 Gyr \\citep{hkn99}, thus the two (WD + MS and WD + RG) channels yield a bimodality of the delay time distribution. (5) The stripped-off material is probably distributed on the orbital plane and forms a massive circumbinary torus (or disk) around SNe Ia. Such circumstellar matter (CSM) may be consistent with the observed CSM feature in SN 2006X. When SN ejecta strongly interact with massive CSM, it can explain the feature of Type Ia/IIn (IIa) SNe 2002ic and 2006gj. (6) Three different environments of SN Ia explosions can be specified by three different states of WDs just at the SN Ia explosion, i.e., the optically thick WD wind phase (Case WIND), steady hydrogen burning phase without optically thick winds from WDs (Case CALM), and recurrent nova phase (Case RN). In Case WIND, SN Ia ejecta strongly interact with massive CSM like SNe Ia/IIn (IIa) 2002ic and 2005gj because CSM exists near the SN Ia. The estimated rate of Case WIND is $\\sim 5$\\% of the total SN Ia rate, being consistent with the observational estimate. In Cases CALM and RN, SNe show a normal SN Ia feature because the CSM is far from the SN but the ejecta may interact with the CSM in a much later phase. SN 2006X may be on a border between Case WIND and Case CALM." }, "0710/0710.5220_arXiv.txt": { "abstract": "{Although many radio loud quasars and galaxies have been observed in X~rays, systematic studies of well defined samples are rare.} {We investigate the X-ray properties of the most luminous radio sources in the 3CR catalogue, in order to assess if they are similar to the most luminous radio quiet quasars, for instance in the X-ray normalization with respect to the optical luminosity, or in the distribution of the absorption column density.} {We have selected the (optically identified) 3CR radio sources whose 178-MHz monochromatic luminosity lies in the highest factor-of-three bin. The 4 most luminous objects had already been observed in X~rays. Of the remaining 16, we observed with XMM-Newton 8 randomly chosen ones, with the only requirement that half were of type~1 and half of type~2 according to the optical identification.} {All targets have been detected. The optical-to-Xray spectral index, $\\alpha_{ox}$, can be computed only for the type~1s and, in agreement with previous studies, is found to be flatter than in radio quiet quasars of similar luminosity. However, the Compton thin type~2s have an absorption corrected X-ray luminosity syste\\-ma\\-ti\\-cally lower than the type~1s, by a factor which makes them consistent with the radio quiet $\\alpha_{ox}$. Within the limited statistics, the Compton thick objects seem to have a reflected component more luminous than the Compton thin ones.} {The extra X-ray component observed in type~1 radio loud quasars is beamed for intrinsic causes, and is not collimated by the absorbing torus as is the case for the (intrinsically isotropic) disk emission. The extra component can be associated with a relativistic outflow, provided that the flow opening angle and the Doppler beaming factor are $\\sim 1/5$ -- $1/7$ radians.} ", "introduction": "A strong X-ray emission is a defining property of Active Galactic Nuclei, and radio loud quasars and galaxies share this property. The most interesting entries of the radio catalogues have been observed repeteadly with all the X-ray satellites launched so far. Although much has been learned, our knowledge is still fragmentary in comparison with the radio quiet AGN. The latter dominate the X-ray sky, and it is relatively straightforward to assemble large samples which can then be studied at other wavelengths. On the contrary, the luminous extragalactic radio sources are rare, and one has to observe them one by one investing large amounts of satellite time. There are a few points on which a consensus has been established. The type~1 radio loud quasars have an X-ray emission stronger than their radio quiet analogues of similar optical luminosity, which is quantified by a flatter $\\alpha_{ox}$ (e.g. \\cite{brink} versus \\cite{steffen}). There are indications that the X-ray photon index $\\Gamma$ is syste\\-ma\\-ti\\-cally flatter in radio loud quasars (\\cite{shastri}, but \\cite {brink} have a more cautious view). At any rate, a flatter $\\Gamma\\sim 1.5$ -- 1 is well established in flat spectrum radio sources and low frequency peaked blazars (\\cite{grandi}, \\cite{fossati}). The type~2 radio galaxies are commonly found to host absorbed AGN, a paradigmatic case is the near\\-by powerful source Cygnus~A (\\cite{young}). Even if absorbed below energies of several keV, the radio galaxies are frequently detected around 1 keV at a level of a few percent of the main emission (\\cite{diana} and references therein). It is unclear if the residual emission is due to scattering of the main one, or is completely unrelated. The basic picture is the extension of the so called unified scheme to the case where the accretion disk is complemented with a relativistic outflow (\"jet\"). The radio galaxies and radio quasars are the misaligned and aligned members, respectively, of the same population. The optical and soft X-ray emission of the former is obscured by some intervening medium, perhaps arranged in a toroidal geometry. The jet emission, because of the Doppler boosting, becomes more and more prominent along a sequence where the line of sight gets closer and closer to the jet axis, passing from steep spectrum to flat spectrum radio quasars, and to blazars. \\begin{table*} \\caption{The sample objects.} \\label{table:1} \\centering \\begin{tabular}{l c c c c c c} \\hline\\hline Name & Type & Redshift & Radio Luminosity & Magnitude & Exposure Time & References \\\\ & & $z$ & L$_{\\rm r}$ & & (ks) & \\\\ \\hline 3C298 & Q & 1.439 & 29.79 & 16.79~(V) & 20~(C) & 1 \\\\ 3C9~~ & Q & 2.012 & 29.72 & 18.21~(V) & 16~(C) & 2 \\\\ 3C257 & G & 2.474 & 29.67 & 18.07~(K) & 30~(X) & 3 \\\\ 3C191 & Q & 1.956 & 29.55 & 18.65~(V) & 17~(C) & 4 \\\\ 3C239 & G & 1.781 & 29.46 & 22.50~(V) & 14~(X) & 5 \\\\ 3C454 & Q & 1.757 & 29.39 & 18.47~(V) & 16~(X) & 5 \\\\ 3C432 & Q & 1.785 & 29.38 & 17.96~(V) & 11~(X) & 5 \\\\ & & & & & 20~(C) & 4 \\\\ 3C294 & G & 1.786 & 29.35 & 18.00~(K) & 118~(C) & 6 \\\\ 3C249 & G & 1.554 & 29.34 & 18.90~(K) & 35~(X) & 5 \\\\ 3C241 & G? & 1.617 & 29.30 & 23.50~(V) & 25~(X) & 5 \\\\ 3C318 & Q? & 1.574 & 29.30 & 20.30~(V) & 19~(X) & 5 \\\\ 3C205 & Q & 1.534 & 29.28 & 17.62~(V) & 5~(X) & 5 \\\\ \\hline \\end{tabular} \\begin{list}{}{} \\item[] References. (1)~\\cite{aneta}; (2)~\\cite{fabian}; (3)~\\cite{derry}; \\item[] (4)~\\cite{erlund}; (5)~our data; (6)~\\cite{fabian2}. \\end{list} \\end{table*} In this paper we present and discuss a study of the X-ray properties of the most luminous low frequency radio sources. By selecting at 178 MHz, the frequency of the 3CR catalogue (\\cite{3cr}), we select sources with a luminous extended, isotropic component, which is thought to arise from the accumulation of relativistic plasma over the entire lifetime of the system. The AGN in our sample are then expected to have had a large {\\it time averaged} power in their past life, whereas the X-ray lu\\-mi\\-no\\-si\\-ty depends on the {\\it instantaneous} power at the present time. A correlation between the two holds only in a statistical sense. On the other hand, by selecting with respect to an isotropic component we avoid all the uncertainties connected with the Doppler boosting. We restrict ourselves to the 3CR sources at high Galactic latitude ($|\\ell|>20^{\\circ}$) with optical identification and redshift (\\cite{ident}), and order them with respect to L$_{\\rm r}$~, a radio luminosity parameter equal to the logarithm of the monochromatic luminosity at 178 MHz in W~Hz$^{-1}$~m$^{-2}$~s$^{-1}$. We adopt the concordance cosmology with $\\Omega_{\\rm m}=0.3, \\Omega_{\\Lambda}=0.7$, and $\\rm H_{\\circ}= 70~km~s^{-1}~Mpc^{-1}$. We neglect the $k$-correction because all the objects of interest happen to be steep spectrum radio sources, i.e. they have a spectral slope $\\alpha$ (f$_{\\nu} \\propto \\nu^{-\\alpha}$) close to 1 in the 100-MHz region. In half a decade, from L$_{\\rm r}$~= 29.79 to L$_{\\rm r}$~=29.28, one has 20 objects. The four most luminous ones had already been observed by previous investigators. We chose at random 8 additional objects among the remaining 16, with the only requirement that 4 were optical type~1 and 4 optical type~2, and observed them with XMM-Newton. One of the 8 (3C294) had previous Chandra observa\\-tions, and we did not reobserve it. One further object (3C432) was also observed with Chandra after our XMM run had been sche\\-du\\-led. Due to the randomness of our choice, we believe that the final sample of 4+8 objects is a fair representation of the highest radio luminosity bin, although a rather slender one. The only infringement to randomness, i.e. the preselection of optical types, implies that we cannot draw inferences about the relative frequency of types from our X-ray data. ", "conclusions": "The main point of the present work has been already anticipated in the previous Section, and is the segregation in intrinsic luminosity between the unabsorbed type~1s and the absorption corrected, Compton thin type~2s. This will be shown to have far reaching implications. However our finding is plagued by the small number of objects involved. In order to put the result on firmer grounds, we have collected all the 3CR sources with X-ray observations having: optical identification and redshift; high Galactic latitude ($|\\ell|>20^{\\circ}$); and radio luminosity parameter L$_{\\rm r}$ in the factor-of-three interval immediately below our fiducial sample, i.e. $ 28.67 < \\rm L_{\\rm r} < 29.28$. The additional sample contains 46 objects, of which only 12 observed in X~rays. With respect to the fiducial sample, the additional one is sparsely observed (26\\% versus 60\\%); furthermore, we have no control on the criteria by which the observed sources were chosen. In order to confirm or disprove the luminosity segregation which we want to investigate, a crucial point is a possible bias of the additional sample with respect to the orientation of the jet axis. One source is the well known blazar 3C454.3; its X-ray emission is completely dominated by the jet, so we do not consider it further in the following. All the remaining 11 sources have steep radio spectra, and none of them is classified as a blazar; for the sake of completeness, we point out that the spectra of three of them (3C380, 3C309.1 and 3C212) become flat above a few GHz, perhaps a hint that their orientation angles are at the lower end of the non-blazar region. Also the distribution of optical properties (4 type 1's, 6 type 2's, and one intermediate type) does not exhibit obvious peculiarities. Table~\\ref{table:3} gives some basic information about the 11 additional sources; more specifically we list: the 3C name; the optical and X-ray type; the redshift $z$; the radio luminosity parameter L$_{\\rm r}$; the intrinsic and reflected X-ray luminosity, L$_{2-10}$ and L$_{\\rm refl}$ (actually, the logarithm of the luminosity in erg s$^{-1}$); and the reference to the X-ray data. Note that 3C325 is a further case of intermediate optical classification (\\cite{grimes}). \\begin{figure}[h!] \\resizebox{\\hsize}{!}{\\includegraphics[angle=0]{salvatifig4nuova.eps}} \\caption{The X-ray luminosity of the sources in our sample (fiducial plus additional), plotted against the radio luminosity parameter. Filled squares, quasars; stars, intermediate optical types; empty squares, Compton thin galaxies (the absorption corrected direct component); filled triangles, Compton thick galaxies; filled triangle between parentheses, 3C249 (see text); empty triangles, Compton thin galaxies (the reflected component). The upper and lower solid lines sketch the regions occupied by type 1's and type 2's, respectively; the dotted lines represent the 5\\% fraction of the solid lines.} \\label{figure:4} \\end{figure} Figure \\ref{figure:4} summarizes the information of Tables \\ref{table:2} and \\ref{table:3}: we plot the logarithm of the X-ray luminosity, Log(L$_{2-10}$) and Log(L$_{\\rm refl}$), against the radio luminosity parameter, L$_{\\rm r}$. The meaning of the symbols is explained in the caption. For the sake of clarity we omit the errorbars: the radio fluxes have typical errors of 1 Jy at the 3C flux limit of 10 Jy, so that the horizontal errorbars are less than 0.04 dex; the X-ray values have typical errors of less than 25\\%, i.e. less than 0.1 dex. Only the reflected components of Compton thin galaxies, the empty triangles in the Figure, have larger errors of the order of 0.3 dex. The lower right corner is expected to be underpopulated, given the limit of 10~Jy for the 3C catalogue and a typical depth of 10$^{-15}$ erg cm$^{-2}$ s$^{-1}$ for the X-ray exposures. The two parallel solid lines sketch the regions occupied by the type~1s and the Compton thin type~2s, respectively. The difference of about 0.75 dex is fully consistent with the average values found in the previous Section for the fiducial sample only, i.e. the addition of more sources confirms the luminosity segregation of the two types. Also, the intermediate optical types are confirmed to lie at intermediate X-ray luminosities. The dotted lines indicate the 5\\% fraction of the type~1 and type~2 luminosity, respectively. This particular value is generally regarded as an upper limit for the reflected or scattered components, because of various constraints on geometry, efficiency, and optical depth. In the case of type~1s the optical emission of the AGN is visible, and one can compute the index $\\alpha_{ox}$. We do that for all the type~1s of the fiducial sample; we derive the rest frame monochromatic fluxes at 2 keV and at 2500 \\AA\\ by means of the observed $\\Gamma$ in the X~rays, and by assuming $\\alpha$ = 0.5 in the optical (f$_{\\nu} \\propto \\nu^{-\\alpha}$). We find $\\alpha_{ox} = 1.35 \\pm 0.11$. At the average optical flux of our sources, 1.9 $10^{31}$ erg cm$^{-2}$ s$^{-1}$ Hz$^{-1}$, the radio quiet AGN have $\\alpha_{ox} = 1.65 \\pm 0.05$ (\\cite{steffen}), thus we recover the well known result (cf. the Introduction) that radio loud quasars are more X-ray luminous than radio quiet ones of the same optical luminosity. This is not true for the radio galaxies, however. If we take into account the lower intrinsic X-ray luminosity of our type~2s we find $\\alpha_{ox} \\sim 1.64$, fully compatible with the radio quiet one. Here we do not observe directly the optical emission of the AGN, and we have assumed that type~2s and type~1s with the same L$_{\\rm r}$ have the same intrinsic optical luminosity. The additional component characteristic of the jetted sources becomes visible at smaller and smaller viewing angles for longer and longer wavelengths; indeed, even the high luminosity blazars exhibit optical broad lines with equivalent widths within a factor of two of the non blazar AGN (\\cite{pian}, \\cite{wills}). This points to the jet optical emission never being dominant over the disk optical emission, and gives some ground to our assumption. The X-ray normalization relative to the optical, and the (meager) evidence provided by the X-ray spectral features (slope $\\Gamma$, equivalent width of the Fe line), suggest that at the large viewing angles typical of radio galaxies we see only the X-ray emission due to an accretion disk very similar to the radio quiet AGN. The self-consistent normalization of X-ray, optical and low frequency radio emission does not leave much room for \"fossil\" quasars, where the extended radio lobes survive for a substantial time after the central engine is switched off. The additional X-ray emission connected with the ``radio loudness'' is not seen in radio galaxies, not because of the intervening absorption, but because this emission is intrinsically beamed. The obvious explanation for the beaming is the relativistic aberration of the jet emission. One can try a more quantitative assessment by adopting the distribution of viewing angles proposed by Barthel (\\cite{bart}), according to whom radio loud AGN appear as type~1s or type~2s if their viewing angle is smaller or larger than $\\sim 44^{\\circ}$; thus the average viewing angle is $\\theta_1\\sim 31^{\\circ}$ and $\\theta_2\\sim 69^{\\circ}$ for the two classes, respectively. Within the type~1 class, one finds the blazar subclass which corresponds to viewing angles smaller than the jet beaming angle, $\\rm \\theta_b < 10^{\\circ}$. If one chooses the most favorable scenario, i.e. a jet of constant length and a very flat spectral slope ($\\Gamma=1$), the ratio of the brightness in the two directions $\\rm \\theta_b$ and $\\theta_1$ is $$ \\rm R_X = [(1-\\beta cos\\theta_1)/(1-\\beta cos\\theta_b)]^2, \\qquad \\beta = \\sqrt{1-\\gamma^{-2}} $$ \\noindent where $\\gamma$ is the bulk Lorentz factor of the jet. This ratio is equal to about 80 for $\\gamma = 15$, as suggested by the analysis of the wide band spectral energy distribution of blazars (\\cite{ghisa}). Indeed, by using the maximum viewing angle for a blazar, $\\rm \\theta_b$, instead of the average one, we have further underestimated $\\rm R_X$. We conclude that the typical X-ray luminosity of high L$_{\\rm r}$ blazars should be $ \\rm R_X$ times the typical X-ray luminosity of steep spectrum quasars of comparable L$_{\\rm r}$, i.e. at least 3 10$^{47}$ erg s$^{-1}$. No such blazar is known all over the sky. A possible solution to this inconsistency could be a geometrically wide, stratified jet: here the narrow and fast spine would be responsible for the blazar properties, while the slower (but still relativistic) sheath would produce the wide angle additional component. Similar schemes have been suggested already, both for AGN (\\cite{chiabe}) and for Gamma Ray Bursts (\\cite{grb}). The above equation for $\\rm R_X$ can in principle be applied to the two directions $\\theta_1$ and $\\theta_2$ to derive the residual jet emission in radio galaxies. This however has nothing to do with the L$_{\\rm refl}$ points in Fig.~\\ref{figure:4}: they refer to a spectral component produced above the absorbing material, at a distance of $\\sim$parsecs from the source, whereas time variability constrains the blazar emission to much smaller dimensions (e.g. \\cite{ghisa}). The correct comparison for the L$_{\\rm refl}$ points is with the 5-percent-fraction dotted lines. One notes that in Fig.~\\ref{figure:4} the empty triangles lie around or below the lower dotted line, which implies that the reflected component of Compton thin type~2s has the expected normalization with respect to the intrinsic disk component. The filled triangles, instead, are systematically higher, and lie around or below the higher dotted line, as if they were normalized to the wide jet component. In other words, if one insisted in attributing the visible emission of Compton thick objects to reprocessing of the disk emission only, one would have to accept reprocessed fractions as high as 30\\%. The statistics is certainly not compelling, however a possible scenario is one where the optical depth at large viewing angles is not the same in all objects. Compton thin and Compton thick type~2 objects would be viewed at similar angles, around $70^{\\circ}$, and would have relatively little and relatively large amounts of matter around them, respectively. Thus the Compton thickness would be related with a larger covered solid angle, perhaps so large as to reprocess also the moderately beamed wide jet emission. Under this hypothesis the peculiar position of 3C241 is nicely accounted for. This source exhibits at the same time an intermediate X-ray luminosity and a relatively large N$_{\\rm H}$, as if the line of sight were grazing at the same time the boundaries of the wide jet and the boundaries of a particularly large absorber: indeed, here the reprocessing fraction is as high as 8\\%, and the corresponding empty triangle falls right in the region of Compton thick sources." }, "0710/0710.2400_arXiv.txt": { "abstract": "Invisible plasma content in blazar jets such as protons and/or thermal electron-positron ($e^{\\pm}$) pairs is explored through combined arguments of dynamical and radiative processes. By comparing physical quantities required by the internal shock model with those obtained through the observed broadband spectra for Mrk 421, we obtain that the ratio of the Lorentz factors of a pair of cold shells resides in about $2\\sim 20$, which implies that the shocks are at most mildly relativistic. Using the obtained Lorentz factors, the total mass density $\\rho$ in the shocked shells is investigated. The upper limit of $\\rho$ is obtained from the condition that thermal bremsstrahlung emission should not exceed the observed $\\gamma$-ray luminosity, whilst the lower limit is constrained from the condition that the energy density of non-thermal electrons is smaller than that of the total plasma. Then we find $\\rho$ is $10^2$-$10^3$ times heavier than that of non-thermal electrons for pure $e^{\\pm}$ pairs, while $10^2$-$10^6$ times heavier for pure electron-proton ($e/p$) content, implying the existence of a large amount of invisible plasma. The origin of the continuous blazar sequence is shortly discussed and we speculate that the total mass density and/or the blending ratio of $e^{\\pm}$ pairs and $e/p$ plasma could be new key quantities for the origin of the sequence. ", "introduction": "The discovery of strong inverse Compton components in $X$ and $\\gamma$-ray emission from jets in active galactic nuclei (hereafter AGN) for a wide range of spatial scales (e.g., Collmar 2001 for review) enables us to probe quantitatively the energetics of relativistic jets. The kinetic power of non-thermal electrons has been estimated by various authors both for inner core jets (i.e., blazars) (e.g., Kino, Takahara and Kusunose 2002, hereafter KTK; Kusunose, Takahara and Kato 2003) and large scale jets (e.g., Tavecchio et al. 2000; Leahy and Gizani 2001, 2002; Kataoka et al. 2003). However, the material content of relativistic jets is not easily constrained by observations since the emission is dominated by that from non-thermal electrons and probably positrons and it is difficult to directly constrain thermal matter content. Hence, the plasma composition in AGN jets, whether normal proton-electron ($e/p$) plasma or electron-positron pairs ($e^{\\pm}$) is a dominant composition, is still a matter of open issue (e.g., Reynolds et al. 1996; Celotti, Kuncic, Rees and Wardle 1998; Wardle et al. 1999; Hirotani et al. 1999; Sikora and Madejski 2000; Ruszkowski and Begelman 2002; Kino and Takahara 2004, hereafter KT04). This problem prevents us from estimating the total mass and energy flux ejected from a central engine. To constrain invisible matter content such as thermal electron-positron pairs and/or protons co-existing with non-thermal electrons, dynamical considerations are indispensable. In KT04, we proposed a new procedure to constrain the invisible thermal plasma component in classical FR II radio sources. We used the fact that the mass and energy densities of the sum of thermal and non-thermal particles are larger than those of non-thermal electrons which are determined by observations. Here we apply the same technique to the inner core jets of AGNs (i.e., blazars) based on the internal shock model. The internal shock model is believed to be most plausible to explain the production of high energy photons and time variabilities in blazars. It has been widely applied also to the prompt emission of gamma-ray bursts (hereafter GRBs) (e.g., Rees 1978; Rees and Meszaros 1994; Kobayashi, Piran and Sari 1997;, Daigne and Mochkovitch 1998; Ghisellini 1999; Spada, Ghisellini, Lazzati and Celotti 2001). It is worth to note that recently Ghisellini et al. (2005) proposed a structured jet model cosisting of a fast spine surrounded by a slowly moving layer for explaining VLBI scale radio blobs. At the present, however, it is not evident where is the acceleration site of electrons in the structured jet model. This is one of the prime issues which should be answered. Internal shocks are potentially the building blocks of the spine part of the structured jet. Whereas we recognize the importance of the detailed structure of jets, as a first step we focus on the physical condition of the flow based on the simple internal shock model. The methodology of constraining the invisible plasma content in the emission region is as follows. As mentioned above, a lower limit to the total mass density (sum of non-thermal electrons and invisible plasma) is restricted by the definition that the mass density of total plasma should be smaller than that of the non-thermal electrons. The mass density of non-thermal electrons can be estimated by multi-frequency observations. For this purpose, in \\S \\ref{sec:shock} we review the shock dynamics of two colliding shells. Note that we do not use the simple two point-mass approximation (e.g., Piran 1999; Lazzati et al. 1999; Zhang and M{\\' e}sz{\\' a}ros 2004 for review) but employ the exact shock dynamics throughout this work. This makes outcomes more accurate. In \\S \\ref{sec:NT} we briefly review the previous results on the amount of non-thermal electrons based on KTK. In \\S \\ref{sec:invisible}, we constrain on the amount of total mass density. As for the upper limit, we use the constraint that bremsstrahlung emission from thermal electron (and positron) component should not exceed the observed $\\gamma$-ray emission. We postulate synchrotron self-Compton (SSC) emission dominance in the $\\gamma$-ray band which is supported by the observed correlations between TeV$\\gamma$-ray and X-ray in TeV blazars (e.g., Takahashi et al. 1996, 2000; Catanese et al. 1997; Maraschi et al. 1999). We can thus bracket the amount of total mass density in the emission region from below and above. In this way we apply this method to the archetypal TeV blazar Mrk 421. In \\S \\ref{sec:dissipation}, we further estimate the shock dissipation rate of the colliding cold shells. The dissipation rate is a widely discussed quantity in literatures concerning gamma-ray bursts (e.g., Lazzati, Ghisellini and Celotti 1999; Piran 1999). The shock dissipation is believed to be the ultimate source of heating and accelerating particles. Summary and discussion are in \\S \\ref{sec:summary}. ", "conclusions": "" }, "0710/0710.0775_arXiv.txt": { "abstract": "The ultraviolet spectra of all ``weak emission line central stars of planetary nebulae'' (WELS) with available {\\it IUE} data are presented and discussed. We performed line identifications, equivalent width and flux measurements for several features in their spectra. We found that the WELS can be divided in three different groups regarding their UV: i) Strong P-Cygni profiles (mainly in \\ion{C}{4} $\\lambda 1549$); ii) Weak P-Cygni features and iii) Absence of P-Cygni profiles. The last group encompasses stars with a featureless UV spectrum or with intense emission lines and a weak continuum, which are most likely of nebular origin. We have measured wind terminal velocities for all objects presenting P-Cygni profiles in \\ion{N}{5} $\\lambda 1238$ and/or \\ion{C}{4} $\\lambda 1549$. The results obtained were compared to the UV data of the two prototype stars of the [WC]-PG 1159 class, namely, A30 and A78. For WELS presenting P-Cygnis, most of the terminal velocities fall in the range $\\sim 1000-1500$ km s$^{-1}$, while [WC]-PG 1159 stars possess much higher values, of $\\sim$3000 km s$^{-1}$. The [WC]-PG1159 stars are characterized by intense, simultaneous P-Cygni emissions in the $\\sim 1150-2000$\\AA\\, interval of \\ion{N}{5} $\\lambda 1238$, \\ion{O}{5} $\\lambda 1371$ and \\ion{C}{4} $\\lambda 1549$. In contrast, we found that \\ion{O}{5} $\\lambda 1371$ is very weak or absent in the WELS spectra. On the basis of the ultraviolet spectra alone, our findings indicate that [WC]-PG 1159 stars are distinct from the WELS, contrary to previous claims in the literature. ", "introduction": "Central stars of planetary nebulae (CSPN) that are hydrogen deficient are generally divided in three main groups: [WR], PG 1159 and [WC]-PG 1159 stars. The first one present spectra with strong and broad emission lines mainly from He, C and O, similar to the Wolf-Rayet stars (WR) of Population I. Depending on the ionization stages of the elements dominating the atmosphere, [WR] stars are subdivided in [WCL], [WCE] and [WO] (Crowther et al. 1998; Acker \\& Neiner 2003). On the other hand, PG 1159 stars are quite distinct objects. They are pre-white dwarfs and show mainly absorption lines of \\ion{He}{2} and \\ion{C}{4} in their spectra (Werner et al. 1997). Only a handful of PG 1159 stars are known to possess wind features (Koesterke et al. 1998; Koesterke \\& Werner 1998). The [WC]-PG 1159 group present strong P-Cygni lines in the ultraviolet (e.g. \\ion{N}{5} $\\lambda 1238$ and \\ion{C}{4} $\\lambda 1549$) and resemble the PG 1159 stars in the optical. Three objects are considered prototypes of this class: A 30, A 78 and Longmore 4 (at outburst). The origin and evolution of hydrogen deficient CSPN have been investigated from different point of views during the last decade. Evolutionary models are now able to provide a reasonable match to the observed chemical abundances, although it is still debated the role of binarity and different thermal pulse models (De Marco \\& Soker 2002; Herwig 2001). Nebular analyses as well as sophisticated non LTE atmosphere models have been also very useful to address several questions regarding these central stars: it is generally inferred that the groups above mentioned form the following evolutionary sequence: late type [WR] $\\rightarrow$ early-type [WR] $\\rightarrow$ [WC]-PG 1159 $\\rightarrow$ PG 1159 $\\rightarrow$ non DA white dwarfs (see e.g. Zijlstra et al. 1994; Pen\\~a et al. 2001; Koesterke 2001). However, this scenario has important issues unsolved, such as the C/He mass ratio and the exact position of [WC]-PG 1159 stars in the HR diagram (Hamann 1997; Marcolino et al. 2007). Moreover, the evolutionary status of the so-called ``weak emission line stars'' and their relation to the [WC]-PG 1159 stars is not at all clear. Let us discuss it now. In an extensive observational study, Tylenda et al. (1993) analyzed spectroscopy data of 77 hydrogen deficient CSPN. In their sample, 39 were classified as [WR] stars. The remaining objects were called ``weak emission line stars'' (WELS or [WELS]). According to these authors, the WELS show emission of \\ion{C}{4} $\\lambda 5805$ (actually \\ion{C}{4} $\\lambda \\lambda 5801, 12$) systematically weaker and narrower than in [WR] stars, and a feature in $\\lambda 4650$, which is possibly a blend of \\ion{N}{3}, \\ion{C}{3} and \\ion{C}{4} emissions. Moreover, \\ion{C}{3} $\\lambda 5696$ is very weak or absent. These optical characteristics were confirmed by Marcolino \\& de Ara\\'ujo (2003) with a homogeneous and higher resolution set of data. Interestingly, by comparing a large sample of WELS, [WC]-PG 1159 and PG 1159 spectra, Parthasarathy et al. (1998) claimed that the WELS and the [WC]-PG 1159 stars actually constitute the same class. However, this claim was based solely on comparisons of optical spectra and was not confirmed by further studies. Indeed, some authors argued that this assertion should be taken with caution until an analysis of a larger sample of objects is performed (e.g. Werner \\& Herwig 2006). Undoubtedly, the ``weak emission line stars'' (WELS) constitute the least understood class of hydrogen deficient CSPN. An important point is that the WELS might not be descendants of the [WR] stars, as it is explicited in the [WR] $\\rightarrow$ [WC]-PG 1159 (=WELS) $\\rightarrow$ PG 1159 evolution. Pen\\~a et al. (2003) for example, derived lower average nebular expansion velocities for WELS than for [WR] stars, while the contrary would be expected from the long-term action of a stellar wind on a planetary nebula during a [WR] $\\rightarrow$ WELS transition. Furthermore, based on a kinematical study of a large sample of planetary nebulae, Gesicki et al. (2006) raised the interesting possibility that some WELS can be progenitors of the hottest [WR] stars, i.e., the [WO] group (WELS $\\rightarrow$ [WO]). A major issue hinders the determination of the evolutionary status of the WELS: their physical parameters (e.g. $T_{eff}$, $v_\\infty$ and $\\dot{M}$) and chemical abundances remain unknown. The properties of A 30, A 78, and Longmore 4 ([WC]-PG 1159 stars) are known (Koesterke 2001) but again, their identity as WELS is questionable. So far, most of previous studies involving WELS were done in the optical part of the spectrum. In fact, the very definition of a WELS is based in the optical features in $\\lambda 4650$, \\ion{C}{4} $\\lambda 5805$ and \\ion{C}{3} $\\lambda 5696$ (Tylenda et al. 1993). A few works in the ultraviolet including some WELS can be found in the literature. However, they often include stars of different spectral classes (e.g. Feibelman 2000), and/or are focused in the determination of nebular properties (e.g. Adams \\& Seaton 1982; Pottasch et al. 2005), without special attention to the WELS and their evolutionary state. Motivated by this fact and considering the open questions above described, we investigate in this paper the UV spectra of all WELS with available {\\it IUE} ({\\it International Ultraviolet Explorer}) data. Our main aims are to understand their main UV characteristics; identify and measure the most intense spectral lines; measure wind terminal velocities from the available P-Cygni profiles; and finally, to compare the results to the data of the two prototype [WC]-PG 1159 stars: A30 and A78. The present paper is organized as follows: in Section 2 we present the observational data retrieved from the Multi-mission Archive at STScI (MAST); in Section 3 we discuss the main characteristics of the UV spectra of the WELS, and present line identifications of the most conspicuous features, as well as equivalent width and line flux measurements. In Section 4 we empirically measure the terminal velocities for all objects presenting P-Cygni profiles in \\ion{N}{5} $\\lambda 1238$ and/or \\ion{C}{4} $\\lambda 1549$, from low and high resolution data. Finally, in Section 5, we discuss and compare the results obtained for the WELS to the data of A 30 and A 78 (the two prototype [WC]-PG 1159 stars), and present the main conclusions of our work. ", "conclusions": "\\label{discussion} From a comparison between several optical spectra of WELS, [WC]-PG 1159, and PG 1159 stars, Parthasarathy et al. (1998) have proposed that the WELS are actually [WC]-PG 1159 stars. This claim was not confirmed by further studies and as we mentioned, some authors have warned about this assertion until a more comprehensive study is achieved (Werner \\& Herwig 2006). After an analysis of the main UV characteristics of the WELS and the two prototype [WC]-PG 1159 stars A30 and A78, our next step was to compare the results obtained for these two class of objects in order to address this important issue. As we have shown, most of the WELS present a UV spectrum considerably different from the [WC]-PG 1159 stars. While this last class present simultaneously P-Cygni profiles in \\ion{N}{5} $\\lambda 1238$, \\ion{O}{5} $\\lambda 1371$, and \\ion{C}{4} $\\lambda 1549$, the majority of WELS present a very weak or no \\ion{O}{5} $\\lambda 1371$ (see Fig. \\ref{group1}). The same is true at least for some objects regarding the \\ion{N}{5} $\\lambda 1238$ line. The only exceptions are the objects NGC 6543, NGC 6567, and NGC 6572. Their spectra in fact resemble the ones of the [WC]-PG 1159 stars (see Figs. \\ref{wcpg} and \\ref{wcpg2}): their \\ion{O}{5} $\\lambda 1371$ line is clearly visible as well as the other transitions mentioned. Besides the spectral differences found in the ultraviolet part of the spectrum, we have also found that the terminal velocities of the WELS are considerably lower than in [WC]-PG 1159 stars. Our Table 6 shows that the bulk of the WELS have $v_\\infty$ between $\\sim 1000-1500$ km s$^{-1}$, and A30 and A78 have values about $3000$ km s$^{-1}$. This difference might represent different physical parameters underlying these two class of stars. The theory of radiatively driven stellar winds for example, predicts that the terminal velocity of a star is related with the escape velocity ($v_{esc}$), which in turn depend on other physical parameters (e.g. mass and radius; Abbott 1978; Lamers \\& Cassinelli 1999). Moreover, $v_\\infty$ is also known to correlate with the effective temperature ($T_{eff}$) in several class of stars (see Fig. 8 of Prinja et al. 1990). The $T_{eff}$ tends to be higher for stars with high terminal velocities. From the considerations above described, we conclude that the [WC]-PG 1159 stars are distinct from the WELS, in contrast with the claim made by Parthasarathy et al. (1998) on the basis of optical spectroscopy. It should be noted however, that the situation for the central stars NGC 6543, NGC 6567 and NGC 6572 is ambiguous. From one side, they have a spectrum compatible with the [WC]-PG 1159 class. On the other hand, they do not present high terminal velocities as it is the case of A30 and A78. From both low and high resolution data it is obtained $v_\\infty$ values less than $2000$ km s$^{-1}$ for these three stars (see Table 6). If the WELS are not [WC]-PG 1159 stars, what is their role in the evolutionary sequence [WR] $\\rightarrow$ PG 1159 ? Do they form an alternative channel of evolution ? In order to elucidate these and other similar questions, and to further clarify the differences between them and the [WC]-PG 1159 stars, we clearly need to determine their physical parameters and chemical abundances. Non LTE expanding atmosphere models are being computed with this purpose by our group with the CMFGEN code of Hillier \\& Miller (1998). In this way, their position in the HR diagram could be determined, a more efficient comparison to [WR], [WC]-PG 1159 and PG 1159 stars could be made, and their evolutionary status could be better determined." }, "0710/0710.2894_arXiv.txt": { "abstract": "In the ongoing HATNet survey we have detected a giant planet, with radius $\\hatcurr\\,\\rjuplong$ and mass $\\hatcurm\\,\\mjuplong$, transiting the bright ($V=10.5$) star \\gschatcur{}. The planet is in a circular orbit with period $\\hatcurP$\\,days and mid-transit epoch \\hatcurT\\,(HJD). The parent star is a late F star with mass $1.29 \\pm 0.06\\,\\msun$, radius $1.46 \\pm 0.06\\,\\rsun$, $\\teff \\sim 6570\\pm80\\,\\mathrm{K}$, $\\mathrm{[Fe/H]} = -0.13 \\pm 0.08$ and age $\\sim 2.3^{+0.5}_{-0.7}\\,\\mathrm{Gy}$. With this radius and mass, \\hatcurb{} has somewhat larger radius than theoretically expected. We describe the observations and their analysis to determine physical properties of the \\hatcur{} system, and briefly discuss some implications of this finding. ", "introduction": "\\label{sec:introduction} The detection of transiting exoplanets is very important to exoplanet research because of the information about both planetary radius and mass that comes from photometric transit light curves combined with follow-up radial velocity observations. The transiting exoplanets known as of this writing span a wide range in the physical parameter space of planetary mass, radius, orbital period, semi-major axis, eccentricity; and parent star parameters, including mass, radius, effective temperature, metallicity, and age. Filling out their distribution in this multidimensional space is certain to give us important information on the origin and evolution of exoplanetary systems. Here we report on the discovery by HATNet of its sixth transiting planet, \\hatcurb{}, an inflated Jupiter-mass gas giant in an essentially circular orbit about an F dwarf star with slightly sub-solar metallicity. ", "conclusions": "\\hatcurb{}, with radius $\\hatcurrshort\\,\\rjup$, is similar in size to five low density ``inflated'' planets tabulated by \\citet{kovacs07} (i.e.~WASP-1b, HAT-P-4b, HD~209458b, TrES-4, and HAT-P-1b). However, its mass of $\\hatcurmshort\\,\\mjup$ is greater than the mass of any of these, and hence it has a larger mean density and surface gravity. For a planet of its mass, age of 2.3 Gy, and stellar flux $F_p$ at the planet given by $F_p = \\lstar / (4 \\pi a^2)$, models of \\cite{Burrows:07} predict a radius of about 1.21 $\\rjup$, assuming that the planet has no heavy-element core. The metallicity of \\hatcur{}, $\\mathrm{[Fe/H]}=-0.13 \\pm 0.08$, is among the smallest of known transiting planet host stars. If we assume that the bulk composition of \\hatcurb{} tracks the metallicity of its host star, the size of its heavy-element core should be small, but not vanishingly so. Thus the predicted radius from \\cite{Burrows:07} is comparable to, but perhaps slightly higher, than one might expect for \\hatcurb{}. However, the actual radius found here, $\\rpl = \\hatcurrlong$, lies above the predicted value by about $2\\sigma$, so it appears to be somewhat inflated relative to that model. \\cite{Hansen:07} proposed that hot jupiters can be placed into two classes based on their equilibrium temperature and Safronov number $\\Theta$, where $\\Theta \\equiv (a/\\rpl) \\times (\\mpl/\\mstar)$ is the ratio of the escape velocity from the surface of the planet to the orbital velocity. When Safronov number is plotted versus equilibrium temperature, transiting hot jupiters seem to fall into two groups, with an absence of objects between. However, the Safronov number for \\hatcurb{} is $0.064 \\pm 0.004$; along with HAT-P-5b \\citep{Bakos:07b} with Safronov number $0.059\\pm0.005$, these two planets appear to fall between the two groups in such a plot. Hansen \\& Barman also noted a difference in the relation between planet mass and equilibrium temperature for planets of the two classes, but HAT-P-6b and HAT-P-5b appear to fall between the two classes in this respect as well. It would seem that discovery and characterization of a large number of additional transiting exoplanets may be necessary to establish unambiguously whether there is a bi-modal distribution of hot jupiter planets according to their Safronov number." }, "0710/0710.4776_arXiv.txt": { "abstract": "The chromosphere of the quiet Sun is a highly intermittent and dynamic phenomenon. Three-dimensional radiation (magneto-)hydrodynamic simulations exhibit a mesh-like pattern of hot shock fronts and cool expanding post-shock regions in the sub-canopy part of the inter-network. This domain might be called ``fluctosphere''. The pattern is produced by propagating shock waves, which are excited at the top of the convection zone and in the photospheric overshoot layer. New high-resolution observations reveal a ubiquitous small-scale pattern of bright structures and dark regions in-between. Although it qualitatively resembles the picture seen in models, more observations -- e.g. with the future ALMA -- are needed for thorough comparisons with present and future models. Quantitative comparisons demand for synthetic intensity maps and spectra for the three-dimensional (magneto-)hydrodynamic simulations. The necessary radiative transfer calculations, which have to take into account deviations from local thermodynamic equilibrium, are computationally very involved so that no reliable results have been produced so far. Until this task becomes feasible, we have to rely on careful qualitative comparisons of simulations and observations. Here we discuss what effects have to be considered for such a comparison. Nevertheless we are now on the verge of assembling a comprehensive picture of the solar chromosphere in inter-network regions as dynamic interplay of shock waves and structuring and guiding magnetic fields. ", "introduction": "The chromosphere of the quiet Sun -- a story full of misunderstandings. Apart from the ongoing controversy concerning the heating mechanism \\cite[(e.g., Fossum \\& Carlsson 2005)]{2005Natur.435..919F)}, many details of the small-scale structure of the chromosphere of inter-network regions are still unknown. Already the term ``chromosphere''\\footnote{\\cite[Rutten (2007, and references therein)]{2007ASPC..368...27R} uses the term ``clapotisphere'' for the shock-dominated subcanopy domain in inter-network regions, whereas his ``chromosphere'' refers to the fibrilar structure visible in H$\\alpha$ only. As ``clapotisphere'' stands for standing waves, we here introduce the term ``fluctosphere'' instead (fluctus = latin for ``wave'').} is a frequent source of misunderstandings . Certainly the large variety of phenomena observed \\cite[(see, e.g., Judge 2006; Rutten 2006, 2007)]{2006ASPC..354..259J,2006ASPC..354..276R,2007ASPC..368...27R} created a complex puzzle and sometimes apparent contradictions. For instance, the observed UV emission implies high temperatures, whereas the existence of carbon monoxide lines point at much cooler gas \\cite[(Ayres 2002)]{ayres02}. New high-resolution observations -- as reported here -- show a highly dynamic and intermittent pattern that cannot be explained with the classical semi-empirical models by \\cite[Vernazza \\etal\\ (1981, VAL)]{val81} and \\cite[Fontenla \\etal\\ (1993, FAL)]{fal93}. Rather a time-dependent three-dimensional model is mandatory. A self-consistent model that can fulfill all observational constraints would be most valuable for summarising the many faces of the chromosphere, indicating and understanding the most relevant processes. Chromospheric heating is a central issue as it has important implications for the atmospheres of other stellar types. Here we report on some advances of detailed radiation magnetohydrodynamic simulations in comparison with new high-resolution observations. Some crucial aspects of such comparisons -- which often result in misunderstandings -- are discussed. ", "conclusions": "State-of-the-art numerical simulations exhibit a highly dynamic chromosphere, which is characterised by propagating and interacting shock waves with co-existing hot and cool regions. New observations, as presented here, now have a sufficiently high spatial, temporal {\\em and} spectral resolution to resolve a pattern of bright structures and dark regions. As synthetic intensity images for the cores of the prominent Ca\\,II lines are still missing, these observations can only be compared to the simulations on a qualitative basis at the moment. Nevertheless they clearly support the picture of the chromosphere as highly dynamic and intermittent phenomenon like it was already implied by the pioneering simulations by \\cite[Carlsson \\& Stein (1995, 1998)]{carlsson95, 1998IAUS..185..435C}. Even \\cite[Kalkofen (2004)]{2004IAUS..219..115K} stated that ``[d]etailed observations show the chromosphere to be highly dynamic''. The model atmospheres by VAL and FAL, although certainly very elaborate, suffer from the basic assumption of a one-dimensional stratified static atmosphere. This assumption is clearly questioned by recent high-resolution observations. Strong intensity fluctuations are now observed although one still should need be careful with deriving statements concerning the gas temperature. Frequently it is argued that the amplitudes and minimum values of gas temperature in the model atmospheres are not observed and that they are wrong as they do not agree with the VAL models \\cite[(cf. Kalkofen 2003b)]{2003SPD....34.1101K}. These arguments obviously do not hold any longer in view of new observational results. VAL-type atmospheres should thus be considered as qualitative averages, which could at best be interpreted as variations on large spatial scales. Instead of aiming at an agreement with VAL models, modern 3D radiation (magneto-)hydrodynamical simulations must be directly compared to observations. Certainly the quantitative values of the temperature fluctuations in these models (apart from the low to middle photosphere) still suffer from the simplified treatment of radiative transfer and resulting uncertainties in the energy balance, which are, however, a necessary compromise in order to keep the computations tractable. The next steps towards realistic chromosphere models requires more work on the modelling of non-equilibrium effects, which are important for the energy balance and thus the temperature amplitudes in the chromosphere. An efficient non-LTE radiative transfer scheme is a major goal. Observational emphasis should be given to (i)~continued aiming at a combination of high spatial \\textit{and} temporal \\textit{and} spectral resolution, as they are crucial for a meaningful comparison and interpretation, and (ii) the development of new diagnostics as, e.g., the (sub-)millimetre continua. Especially the upcoming ALMA might allow a tomography of the solar atmosphere, finally revealing details of its three-dimensional structure." }, "0710/0710.3683_arXiv.txt": { "abstract": "We analyze the rotation curves of 10 spiral galaxies with a newtonian potential corrected with an extra logarithmic term, using a disc modelization for the spiral galaxies. There is a new constant associated with the extra term in the potential. The rotation curve of the chosen sample of spiral galaxies is well reproduced for a given range of the new constant. It is argued that this correction can have its origin from string configurations. The compatibility of this correction with local physics is discussed. ", "introduction": " ", "conclusions": "" }, "0710/0710.3010_arXiv.txt": { "abstract": "Recent observations of the Mira AB binary system have revealed a surrounding arc-like structure and a stream of material stretching 2 degrees away in opposition to the arc. The alignment of the proper motion vector and the arc-like structure shows the structures to be a bow shock and accompanying tail. We have successfully hydrodynamically modelled the bow shock and tail as the interaction between the asymptotic giant branch (AGB) wind launched from Mira A and the surrounding interstellar medium. Our simulations show that the wake behind the bow shock is turbulent: this forms periodic density variations in the tail similar to those observed. We investigate the possiblity of mass-loss variations, but find that these have limited effect on the tail structure. The tail is estimated to be approximately 450\\,000 years old, and is moving with a velocity close to that of Mira itself. We suggest that the duration of the high mass-loss phase on the AGB may have been underestimated. Finally, both the tail curvature and the rebrightening at large distance can be qualitatively understood if Mira recently entered the Local Bubble. This is estimated to have occured 17\\,pc downstream from its current location. ", "introduction": "In the Mira binary system, Mira A is an evolved star which is undergoing a period of enhanced mass-loss as it moves along the asymptotic giant branch (AGB) on route to becoming a white dwarf. The companion star, previously classified as a white dwarf, now has a less clear classification \\citep{karovska05,ireland07} but is less luminous and any stellar outflow is comparably insignificant to Mira A in terms of mass-flux and energetics. New UV observations of the Mira system \\citep{martin07} have revealed a comet-like tail extending 2 degrees away to the North and an arc-like structure in the South. \\cite{martin07} postulated that these features are a bow shock and a ram-pressure-stripped tail stretching away in opposition to the bow shock caused by motion through the interstellar medium (ISM). Their postulation is consistent with Mira's proper motion \\citep{turon93} of 225.8 milli-arcseonds per year in the direction 187.1 degrees East of North (corrected for solar motion). Further, they note that at the revised Hipparcos-based distance \\citep{knapp03} of 107 pc, the large space velocity of 130 \\kms, calculated from the proper motion and the radial velocity \\citep{evans67} of 63 \\kms, is further consistent with the bow shock structure. In this paper, we present hydrodynamical modelling of the Mira system and discuss the implications with respect to postulation of \\cite{martin07}. We are aiming to fit the position of the bow shock ahead of the central star, the width across the central star, the undulating density profile along the tail, the length of the tail and the ring-like structure one third the way down the tail. We show the GALEX observation in Figure 1. ", "conclusions": "Our models for Mira's evolution suggest that the tail traces half a million years of mass-loss history and ISM interaction. An instability at the bow shock is the cause of the fluctuations seen in the tail. Mass-loss variations of Mira A have much less effect. The curvature in the tail is caused by the differential velocity between Mira and its tail leading to different Galactic orbits. We attribute a gap in the tail to Mira entering the local bubble, 17 pc downstream from its current location. The time to reestablish the bow shock directly leads to the gap in the tail. A ring-like structure is attributed to a vortex shed into the tail just before the bow shock reached its equilibrium position. The tail therefore traces not only the history of Mira itself, but also the structure of the ISM along its path. It's a wonderful prospect." }, "0710/0710.5636_arXiv.txt": { "abstract": "We present a parametric strong lensing model of the cluster Abell 2218 based on HST ACS data. We constrain the lens model using 8 previously known multiply imaged systems, of which 7 have spectroscopically confirmed redshifts. In addition, we propose five candidate multiply imaged systems and estimate their redshifts using our mass model. The model parameters are optimized in the source plane by a bayesian Monte Carlo Markov Chain as implemented in the the publicly available software Lenstool. We find rms$_s=0\\farcs12$ for the scatter of the sources in the source plane, which translates into rms$_i=1\\farcs49$ between the predicted and measured image positions in the image plane. We find that the projected mass distribution of Abell 2218 is bimodal, which is supported by an analysis of the light distribution. We also find evidence for two structures in velocity space, separated by $\\sim 1000$~km~$s^{-1}$, corresponding to the two large scale dark matter clumps. We find that the lensing constraints can not be well reproduced using only dark matter halos associated with the cluster galaxies, but that the dark matter is required to be smoothly distributed in large scale halos. At $100\\arcsec$ ($291$~kpc) the enclosed projected mass is $3.8\\times10^{14}$~M$_\\sun$. At that radius, the large scale halos contribute $\\sim85\\%$ of the mass, the brightest central galaxy (BCG) contributes $\\sim9\\%$ while the remaining $\\sim6\\%$ come from the other cluster galaxies. We find that the model is not very sensitive to the fainter (and therefore by assumption less massive) galaxy sized halos, unless they locally perturb a given multiply imaged system. Therefore, dark galaxy sized substructure can be reliably constrained only if it locally perturbs one of the systems. The massive BCG and galaxies which locally perturb a multiply imaged system are reliably detected in the mass reconstruction. In an appendix we give a self-contained description of the parametric profile we use, the dual pseudo isothermal elliptical mass distribution (dPIE). This profile is a two component pseudo isothermal mass distribution (PIEMD) with both a core radii and a scale radii. ", "introduction": "Dark matter dominates over baryonic matter in the universe, but its nature is not known. The study of the inner parts of dark matter halos can give insight into the nature of the dark matter, as the steepness of the profile is correlated with the interaction between the dark matter itself and with the baryonic matter. According to $\\Lambda$CDM simulations, the mass distribution of galaxy clusters should be dominated by their dark matter halos. Gravitational lensing, which is sensitive to the total matter distribution, visible or dark, is ideal for studying the mass distribution of clusters. Strong lensing features, consisting of multiply imaged and strongly distorted background sources, provide constraints on the inner parts of the cluster, while weak lensing features, consisting of weakly distorted singly imaged background sources, provide constraints on the outer slope of the surface mass profile \\citep[see e.g.,][]{smail1995a,smail1995b,kneib1996, smail1997,abdelsalam1998,natarajan2002b, bradac2006,gavazzi2007,limousin2007}. Lensing can therefore provide unique information about the total mass distribution of clusters, from the inner to the outer parts. In addition, lensing can in principle be used to deduce various cosmological parameters (e.g. $H_0$, $\\Omega_\\Lambda$, $\\Omega_m$). This has already been extensively applied to lensing on galaxy scales \\citep[see e.g.,][]{schechter1997,koopmans2003} and to a smaller degree for lensing on cluster scales \\citep[see e.g.,][]{soucail2004, meneghetti2005}, where the accuracy of the mass map is a limiting factor. The accuracy of the mass map is strongly dependent on the number of multiply imaged systems used to constrain it. Therefore, to construct a robust model of the dark matter distribution, accurate enough for cosmography and for using the cluster as a gravitational telescope, it is important to include as many spectroscopically confirmed multiply imaged systems as possible \\citep[see e.g.,][]{ellis2001}. Abell 2218 is one of the richest clusters in the Abell galaxy cluster catalog \\citep{abell1958, abell1989} and has been successfully exploited as a gravitational lens. A parametric lens model has previously been constructed by \\citet{kneib1995,kneib1996} (using 1 and 2 spectroscopically confirmed systems respectively) and by \\citet{natarajan2002b, natarajan2007} (using 4 and 5 spectroscopically confirmed systems) building on the model of \\citet{kneib1996} and including weak lensing constraints from HST WFPC2 data. A non-parametric model was constructed by \\citet{abdelsalam1998} using three spectroscopically confirmed multiply imaged systems. In all of these models, a bimodal mass distribution was required to explain the image configurations (i.e. the models include two large scale dark matter clumps), but the number of constraints were not sufficient to accurately constrain the second large scale dark matter clump. Abell 2218 has also been used as a gravitational telescope, with \\citet{ellis2001} discovering a source at $z=5.6$ and \\citet{kneib2004} discovering an even more distant source at $z\\sim6.7$, later confirmed by \\citet{egami2005} using Spitzer data. \\citet{soucail2004} estimated cosmological parameters based on a lensing model of Abell 2218 using 4 multiply imaged systems. The latest published lens model of Abell 2218 is by \\citet{smith2005}, who incorporated four multiply imaged systems and weak lensing constraints, using the WFPC2 data. Although the number of constraints has increased from the initial models, all previous models have assumed that the location of the second dark matter clump coincided with the brightest galaxy in the South East, due to a lack of constraints in its vicinity. The motivation for revisiting the modeling of Abell 2218 comes from the new ACS data which have not been used before for modeling this cluster and are superior in both resolution, sensitivity and field of view to the previous WFPC2 data set. These new high quality data allow us to identify several subcomponents in previously known multiple images, thus adding more constraints, and in one case, two more multiple images of a known system. In addition, we have measured a spectroscopic redshift for an arc around the second dark matter clump, which our model predicts to be singly imaged. We also have several new candidate multiply imaged systems, which we add as constraints and estimate their redshift with the model. In total we have 7 multiply imaged systems with measured spectroscopic redshift and 6 multiply imaged systems without spectroscopic data (of which 5 are new candidate systems). Finally, the lensing code Lenstool, has undergone significant improvements from previous models, with the incorporation of a Monte Carlo Markov Chain (MCMC), which enables us to not only find the best model in the lowest $\\chi^2$ sense, but the most likely model as measured by its Evidence \\citep{jullo2007}. The MCMC also allows for a reliable estimate of the uncertainties in the derived model parameters. The paper is organized as follows: In Section~\\ref{sec:data}, we give an overview of the data used in this paper. We compile a list of all currently known and new multiply imaged systems in Abell 2218 and discuss the reliability of the redshift estimate of each one in Section~\\ref{sec:systems}. The methodology of the strong lensing modeling is described in Section~\\ref{sec:modeling}. In Section~\\ref{sec:analysis} we present the results of our lensing analysis, and compare them to previous models. In Section~\\ref{sec:degeneracy} we discuss degeneracies in the modeling. In Section~\\ref{sec:reliability} we address how reliable our model is, and discuss the smoothness of the dark matter distribution. In Section~\\ref{sec:bimodal} we interpret the bimodality of our model, along with X-ray measurements and an analysis of the distribution of cluster members in velocity space. We summarize our main conclusions in Section~\\ref{sec:conclusions}. Throughout the paper, we adopt a flat $\\Lambda$-dominated Universe with $\\Omega_\\Lambda=0.7$, $\\Omega_m=0.3$ and $H_{0}=70\\ \\mathrm{km\\,s}^{-1}\\,\\mathrm{Mpc}^{-1}$. Following \\citet{smith2005} we place the cluster at $z=0.171$. At this redshift $1\\arcsec$ corresponds to $2.91$~kpc in the given cosmology. ", "conclusions": "" }, "0710/0710.5770_arXiv.txt": { "abstract": "We investigate the distribution of massive black holes (MBHs) in the Virgo cluster. Observations suggest that AGN activity is widespread in massive galaxies ($M_*\\simgt 10^{10}\\msun$), while at lower galaxy masses star clusters are more abundant, which might imply a limited presence of central black holes in these galaxy-mass regimes. We explore if this possible threshold in MBH hosting, is linked to {\\it nature}, {\\it nurture}, or a mixture of both. The {\\it nature} scenario arises naturally in hierarchical cosmologies, as MBH formation mechanisms typically are efficient in biased systems, which would later evolve into massive galaxies. {\\it Nurture}, in the guise of MBH ejections following MBH mergers, provides an additional mechanism that is more effective for low mass, satellite galaxies. The combination of inefficient formation, and lower retention of MBHs, leads to the natural explanation of the distribution of compact massive objects in Virgo galaxies. If MBHs arrive to the correlation with the host mass and velocity dispersion during merger-triggered accretion episodes, sustained tidal stripping of the host galaxies creates a population of MBHs which lie above the expected scaling between the holes and their host mass, suggesting a possible environmental dependence. ", "introduction": "The nearby Virgo cluster is a perfect laboratory to investigate the evolution of galaxies in a dense environment. Recently, observations of a large sample of galaxies suggested that the properties of nuclei, either quiescent or active, in Virgo galaxies are strongly mass-dependent. This latter finding is in very good agreement with the general trend found also in the SDSS, where very few AGN are found in galaxies with stellar mass $M_*< 10^{10}\\msun$\\citep{Kauffmann2003, Kewley2006}. Decarli et al. (2007) analyzed nuclear activity in late type galaxies in the Virgo cluster. They conclude, quite remarkably, that at galaxy mass\\footnote{Decarli et al. (2007) measure the dynamical mass of the galaxy within the optical radius, determined at the $25^{th}$ mag arcsec$^{-2}$ isophote in the B band. This is an upper limit to the stellar mass of galaxies, but a lower limit to the total baryonic mass.} $M_{gal}> 10^{10.5}\\msun$ the AGN fraction is unity. As a central black hole is a necessary condition for AGN activity, we conclude that the black hole occupation fraction (BHOF) must be unity as well. \\cite{Cote2006}, Wehner \\& Harris (2006) and Ferrarese et al. (2006) find that, below $M_*\\sim 10^{10}\\msun$, Virgo galaxies exhibit nuclear star clusters, whose mass scales with $M_*$ in the same fashion as those of the massive black holes detected in brighter galaxies \\citep{Magorrian1998, MarconiHunt2003, Haring2004}. Although the existence of a nuclear star cluster does not rule out a small, hidden MBH, it is suggestive that Ferrarese et al. (2006) conclude that ``bright galaxies often, and perhaps always, contain supermassive black holes but not stellar nuclei. As one moves to fainter galaxies, nuclei become the dominant feature while MBHs might become less common and perhaps disappear entirely at the faint end.\" There are three interlaced sides of the intriguing story which appears to link stellar nuclei and MBHs: 1) understand if (and why) MBHs populate preferentially bright galaxies; 2) understand why stellar nuclei populate preferentially faint galaxies and 3) understand why the ratio of nuclear to galaxy mass is identical to the ratio of MBH to galaxy mass. In this paper we address the first issue. A possible hint to explain the predominance of star clusters in small galaxies may come from comparing the dynamical time scale and the fragmentation time scale of the infalling gas. Detailed calculations are necessary to test this hypothesis, but it can be argued that in shallow potentials gas could fragment before reaching the center of the galaxy. This is even more suggestive in the case of merger-induced gaseous infall. Regarding the third issue, \\cite{Emsellem2007} for instance show that if galaxies have Sersic profiles, radial compressive forces trigger the collapse of gas in the central regions, and the mass of the nuclear cluster that forms is about 0.1\\% of the mass of the host galaxy. So, nuclear cluster formation and MBH feedback might produce the same scaling relation with galaxy mass, but it is unclear whether this is a coincidence or the result of a single, unexplored process. Volonteri et al. (2007) investigate the overall distribution of MBHs in galaxies, as a function of the host velocity dispersion, $\\sigma_*$ , as traced by the host halo (cfr. Ferrarese 2002), and find that the efficiency of MBH formation decreases with halo mass, and isolated dwarf galaxies are most likely to be devoid of a central MBH. This is a common feature of MBH formation models which invoke gas-dynamical processes in the high redshift Universe, either via direct collapse \\citep[e.g.,]{haehnelt1993,LoebRasio1994,Eisenstein1995,BrommLoeb2003,BegelmanVolonteriRees2006,LN2006}, or via an intermediate stage of metal free Population III star \\citep[e.g.,]{CBA84,MadauRees2001,VHM}. The common feature is the need for deep potential wells at cosmic epochs when the protogalaxy population was dominated by minihalos ($M_h<10^7\\msun$). \\\\ \\indent A second physical phenomenon strengthens the likelihood that dwarf galaxies, even if seeded with a MBH at early times, are now lacking one: the gravitational recoil. That is, the general-relativistic effect which imparts velocity to the center of mass of a merging MBH binary, when the MBHs plunge into each other's last stable orbit by emission of gravitational radiation. Gravitational waves carry, in general, a non-zero net linear momentum, which establishes a preferential direction for the propagation of the waves. As a consequence, the center of mass of the binary recoils in the opposite direction \\citep{redmountrees}, possibly causing the ejection of MBHs from the potential wells of their host galaxies \\citep[e.g.,][]{Madau2004,MadauQuataert2004, Merrittetal2004, VolonteriRees2006, Haiman2004, SchnittmanBuonanno2007}. The progress of numerical relativity is now leading to a convergence in the estimates of the recoil. Schwarzschild, i.e., non-spinning, black holes \\citep[e.g.,][]{Bakeretal2006} are expected to recoil with velocities below 200 $\\rm{km\\,s^{-1}}$, and a similar range is expected for black holes with low spins, or with spins (anti-)aligned with the orbital axis. However, when the spin vectors have opposite directions and are in the orbital plane, the recoil velocity can be as large as a few thousands $\\rm{km\\,s^{-1}}$ \\citep{Campanelli2007b,Gonzalez2007,Campanelli2007}. Schnittman (2007) shows that the hierarchical nature of galaxy assembly implies that ejections do not lead to void nuclei, even in the high-recoil limit. This is weaved into the very nature of galaxy assembly, via a series of mergers, in a pattern typically dubbed ``merger tree\". The botanical essence of the tree implies that many branches converge into a central trunk, so that in every generation of the tree, the number of galaxies decreases, while the fraction of MBHs can increase, even if ejections operate. This is especially true for ``central galaxies\", which represents the trunk of the merger tree. The situation is different for satellites in a galaxy cluster: they correspond to loose branches that do not merge into the main trunk. Satellites are indeed galaxies that enter the dark matter halo of the cluster, but do not merge with the central galaxy. In this paper we develop simple models that describe the dynamical evolution of satellite galaxies in a cluster, focusing on the fate of their nuclei. We estimate the influence of MBH formation mechanisms, and of MBH ejections for shaping the BHOF in the Virgo cluster today. ", "conclusions": "We explored the influence of formation epoch of MBHs, bias of their hosts, and MBHs dynamical evolution on the occupation distribution of MBHs in galaxy clusters. The possible decrease of the occupation fraction at low galaxy mass proposed by \\cite{Wehner2006} and \\cite{Ferrareseetal2006} is not a surprising result and follows naturally from the evolution of the MBH population of galaxies in clusters. \\begin{itemize} \\item When the formation mechanisms of MBH seeds are taken into consideration, implying formation of MBHs in massive high redshift halos, the BHOF in cluster galaxies in an increasing function of galaxy mass, in line with observations of the AGN fraction in Virgo. \\item The exact mass threshold above which BHOF=1 depends on the details of the formation mechanism (host masses and redshift of formation, ``nature\") and on the dynamical evolution, including MBH spin magnitude (``nurture\"). \\item The repercussions of ``nurture\" are magnified in cluster galaxies. If a satellite galaxy looses its MBH due to a dynamical interaction, it has a negligible chance of capturing a new one following a subsequent galaxy merger, except for the central galaxy in the cluster. \\item We also predict that if MBHs co-evolve with galaxies during galaxy mergers, and satellite galaxies experience tidal stripping during their orbital evolution, then a population of galaxies where the MBH mass lies above the standard $M_{BH}-M_{gal}$ is expected. Two such examples could be NGC~4486B and NGC~4342. \\end{itemize} We emphasize that our models do not in any way imply that MBHs and nuclear clusters are mutually exclusive. We however predict that the occupation fraction decreases with galaxy mass. If, as suggested by Wehner \\& Harris (2006), lacking a MBH is a necessary condition for nuclear cluster formation, radio or hard X-ray observations will inform us of the complementary or mutually exclusive essence of MBHs and nuclear clusters. \\cite{Gallo2007} recently observed 32 galaxies in Virgo with the Chandra X-ray Observatory, and found, in agreement with \\cite{Decarli2007} that nuclear X-ray activity increases with the mass of the host galaxy. Intriguingly, at least in one case, VCC1178, a nuclear X-ray source is detected jointly with a central star cluster. \\cite{Seth2006} also report that a small AGN might be hosted within the core of the nuclear star cluster in NGC 4206." }, "0710/0710.2199_arXiv.txt": { "abstract": "We report on active region EUV dynamic events observed simultaneously at high-cadence with SUMER/SoHO and TRACE. Although the features appear in the TRACE~Fe~{\\sc ix/x}~171~\\AA\\ images as jets seen in projection on the solar disk, the SUMER spectral line profiles suggest that the plasma has been driven along a curved large scale magnetic structure, a pre-existing loop. The SUMER observations were carried out in spectral lines covering a large temperature range from 10$^4$~K to 10$^6$~K. The spectral analysis revealed that a sudden heating from an energy deposition is followed by a high velocity plasma flow. The Doppler velocities were found to be in the range from 90 to 160 \\kms. The heating process has a duration which is below the SUMER exposure time of 25 {\\rm s} while the lifetime of the events is from 5 to 15 {\\rm min}. The additional check on soft X-ray Yohkoh images shows that the features most probably reach 3 MK (X-ray) temperatures. The spectroscopic analysis showed no existence of cold material during the events. ", "introduction": "A large variety of jet-like phenomena are often observed in the solar atmosphere such as surges, spicules, sprays, Extreme-UltraViolet (EUV) and X-ray jets. X-ray jets (Shibata \\etal\\ 1992) were first identified in data obtained with the Soft X-ray Telescope (SXT) on Yohkoh (Tsuneta \\etal\\ 1991). They represent X-ray enhancements with an apparent collimated motion and were found to have a typical size of 5~$\\times$~10$^3$ -- 4~$\\times$~10$^5$~{\\rm km} and an apparent velocity of 30 to 300 \\kms. Their kinetic energy is estimated to be 10$^{25}$ -- 10$^{28}$~{\\rm ergs}. Most of the jets were associated with small flares in large X-ray bright points or active regions. Shimojo \\& Shibata (2000) derived the physical parameters of X-ray jets and found temperatures from 3 to 8 MK (determined by using Yohkoh filter ratios) and densities of 0.7 -- 4.0 $\\times$ 10$^9$~{\\rm cm}$^{-3}$. It is strongly believed that they are produced by magnetic reconnection and represent the evaporation flow resulting from the reconnection heating. EUV jets were studied by Brekke (1999) in off-limb data from the Coronal Diagnostics Spectrometer (CDS) and the Extreme-ultraviolet Imaging Telescope (EIT). From the CDS data it was found that the jet was emitting only at transition region temperatures showing Doppler shifts in the O~{\\sc v} 629.73 \\AA\\ line up to -75 \\kms. The event was also seen in the EIT Fe~{\\sc xii}~195~\\AA\\ passband propagating with an apparent velocity of 180 \\kms. The plasma seemed to be ejected along a large looped magnetic structure. Jets were also analysed in on-disk data from the Transition Region And Coronal Explorer (TRACE) taken in the 171~\\AA\\ and 1216~\\AA\\ passbands by Alexander \\& Fletcher (2000). In the 171~\\AA\\ channel the ejected plasma was seen both in emission and absorption which suggests that simultaneously highly collimated hot and cold material was ejected along the magnetic field lines. An EUV jet from a new emerging active region (a large Bright Point) was analysed in simultaneous TRACE, EIT and CDS data by Lin \\etal\\ (2006). The authors found the plasma jet to emit in a wide temperature range from 10\\,000~K (He~{\\sc i}) to 2.5~MK (Fe~{\\sc xvi}, the upper temperature limit of their observations). H$_\\alpha$ surges are often associated with EUV and X-ray emissions showing the co-existence of cool (H$_\\alpha$) and hot ejections of plasma (Jiang \\etal\\ 2007 and references therein). Only recently, however, have the spatial and temporal relation of these emissions been studied in detail (Jian \\etal\\ 2007) during surge events in a plage area of an active region. The authors first observed the bright structures in TRACE 171 \\AA\\ followed by the cooler H$_\\alpha$ jet which they interpret as cooling of the hot plasma with a cooling time lasting about 6 -- 15 {\\rm min}. ", "conclusions": "This letter presents, to our knowledge for the first time, EUV transient features in an active region identified and analysed in on-disk SUMER data and simultaneously obtained TRACE images. These instruments provide data at highest existing (1.5\\arcsec and 1\\arcsec) spatial and 2 \\kms\\ spectral resolution (SUMER). Additionally, the combination of spectrometer and imager data obtained at high cadence (25~{\\rm s} and below) permitted the temporal and spatial evolution, velocities and especially temperatures of EUV active region transients to be derived with the highest possible precision. Three dynamic events were studied in spectral lines covering a temperature range from 10$^4$ to 10$^6$~K. All three features showed strong red-shifted emission in the O~{\\sc v}~629~\\AA\\ and Mg~{\\sc x}~625~\\AA\\ lines suggesting high velocity flows which propagate in a direction away from the observer, i.e. towards the solar surface. Considering the magnetic fields structure of the active region field by the loops seen in TRACE~171 \\AA, we suggest that the features although appearing with a jet-like structure in TRACE~171 \\AA\\ images may rather represent a high velocity flow driven along a curved magnetic field, most probably a pre-existing loop. No signature of the events was found at chromospheric temperatures. Both EIT/Fe~{\\sc xii}~195~\\AA\\ and Yohkoh/SXT showed brightenings in a pixel row indicating the presence of a 1 to 3 MK plasma during the transients. The lower resolution of these instuments ($\\approx$6\\arcsec) in comparison to TRACE (1\\arcsec) do not permit the events to be identified as jets. The response in the transition region lines is delayed by around 2~{\\rm min} in respect to the coronal line suggesting cooling of the events. In the future hydrodynamic numerical simulations (see for detail Doyle \\etal\\ 2006b) will be performed with the results converted into observable quantities to be then compared with the present data. We believe that the capabilities of the Hinode mission will bring a better understanding on these features and, more important, the physical mechanism behind them." }, "0710/0710.0964_arXiv.txt": { "abstract": "We use a large suite of carefully controlled full hydrodynamic simulations to study the ram pressure stripping of the hot gaseous halos of galaxies as they fall into massive groups and clusters. The sensitivity of the results to the orbit, total galaxy mass, and galaxy structural properties is explored. For typical structural and orbital parameters, we find that $\\sim 30\\%$ of the initial hot galactic halo gas can remain in place after 10 Gyr. We propose a physically simple analytic model that describes the stripping seen in the simulations remarkably well. The model is analogous to the original formulation of Gunn \\& Gott (1972), except that it is appropriate for the case of a spherical (hot) gas distribution (as opposed to a face-on cold disk) and takes into account that stripping is not instantaneous but occurs on a characteristic timescale. The model reproduces the results of the simulations to within $\\approx 10\\%$ at almost all times for all the orbits, mass ratios, and galaxy structural properties we have explored. The one exception involves unlikely systems where the orbit of the galaxy is highly non-radial and its mass exceeds about 10\\% of the group or cluster into which it is falling (in which case the model under-predicts the stripping following pericentric passage). The proposed model has several interesting applications, including modelling the ram pressure stripping of both observed and cosmologically-simulated galaxies and as a way to improve current semi-analytic models of galaxy formation. One immediate consequence is that the colours and morphologies of satellite galaxies in groups and clusters will differ significantly from those predicted with the standard assumption of complete stripping of the hot coronae. ", "introduction": "There are marked differences in the observed properties of the field and cluster galaxy populations. Perhaps the best known difference is the larger fraction of galaxies that are ellipticals or S$0$s (and the correspondingly lower spiral fraction) in clusters relative to the field (e.g., Dressler 1980; Goto et al.\\ 2003). Not only are the morphologies of cluster galaxies different from those of field galaxies, but so too are a variety of their other observed properties, including colours (e.g., Balogh et al.\\ 2004; Hogg et al.\\ 2004), star forming properties (e.g., Poggianti et al. 1999; Balogh et al.\\ 2000; Gomez et al.\\ 2003), and the distribution and total mass of their gaseous component (e.g., Cayatte et al.\\ 1994; Solanes et al.\\ 2001). These observed differences indicate that the dense environments of groups and clusters are somehow strongly modifying the properties of galaxies as they fall in. Uncovering the physical mechanisms that give rise to the observed variation in galaxy properties has been an active topic of research over the past two or three decades (e.g., Dressler 1984; Sarazin 1988). One of the most commonly mentioned processes is ram pressure stripping (Gunn \\& Gott 1972). Here the gaseous component (which can be composed of both cold atomic/molecular gas and a hot extended component) of the orbiting galaxy is subjected to a wind due to its motion relative to the intracluster medium (ICM). The gas will be stripped if the wind is sufficiently strong to overcome the gravity of the galaxy. Recently, direct observational evidence for the ram pressure stripping of galaxies in clusters has been provided by long (up to tens of kpc) tails of gas seen to be trailing behind several cluster galaxies (e.g., Sakelliou et al.\\ 2005; Crowl et al.\\ 2005; Vollmer et al.\\ 2005; Sun \\& Vikhlinin 2005; Machacek et al.\\ 2006; Sun et al.\\ 2007a). Such stripping could at least partially account for the differing properties of cluster and field galaxies. There have been numerous theoretical studies dedicated to calculating the effects of ram pressure stripping on galaxies using hydrodynamic simulations or semi-analytic models. The vast majority of these studies have focused on the stripping of cold gaseous disks with an emphasis on whether this can account for the observed lower star formation rates (and redder colours) of cluster spirals relative to their field counterparts (e.g., Abadi et al.\\ 1999; Quilis et al.\\ 2000; Vollmer et al.\\ 2001; Okamoto \\& Nagashima 2003; Mayer et al.\\ 2006; Roediger et al.\\ 2006; Hester 2006; Jachym et al.\\ 2007; Roediger \\& Br{\\\"u}ggen 2006; 2007). However, the stripping of extended {\\it hot} gaseous halos of galaxies is only just beginning to be explored (e.g., Kawata \\& Mulchaey 2007) and has not yet been studied in a detailed and systematic way. The hot extended component is predicted to exist around most massive galaxies by semi-analytic models and cosmological simulations and is directly observable at X-ray wavelengths in the case of normal ellipticals. If the hot gaseous halo is completely stripped (as is typically assumed), the only fuel available for star formation is that which resided in the cold component when the galaxy first fell into the cluster. (This process of removing the supply of halo gas is sometimes referred to as ``strangulation'' or ``starvation''.) However, if the hot halo remains intact for some time it can, via radiative cooling losses, replenish the cold component and potentially significantly prolong star formation. This, in turn, would affect the colours and morphologies of cluster galaxies (e.g., Larson et al.\\ 1980; Abadi et al.\\ 1999; Benson et al.\\ 2000; Balogh et al.\\ 2000). Aspects of the stripping of the {\\it hot} gaseous halos of galaxies in clusters have been considered in previous work (e.g., Gisler 1976; Sarazin 1979; Takeda et al.\\ 1984). Mori \\& Burkert (2000) studied the stripping of dwarf galaxies subject to a constant wind using two-dimensional simulations and found that the relatively shallow potential wells of these systems cannot retain their hot gas component for long. However, these authors did not study more massive systems, such as normal ellipticals and spirals, where stripping of the hot ($\\ga 10^6$ K) halo should be much less efficient due to their higher masses and deeper potential wells. [Indeed, a recent X-ray survey of massive galaxies in hot clusters by Sun et al.\\ (2007b) has revealed that {\\it most} of the galaxies have detectable hot gaseous halos.] A few other studies have examined the stripping of more massive systems but not in the context described above. In particular, they have largely focused on the metal enrichment of the ICM (e.g., Schindler et al.\\ 2005; Kapferer et al.\\ 2007), the X-ray properties of the galaxies (Toniazzo \\& Schindler 2001; Acreman et al.\\ 2003) or the generation of ``cold fronts'' (e.g., Takizawa 2005; Ascasibar \\& Markevitch 2006). In the present paper, we carry out a detailed study of the ram pressure stripping of the hot gaseous halos of galaxies as they fall into groups and clusters. This is performed using a large suite of controlled hydrodynamic simulations. Unlike most previous studies, we use full three-dimensional (3D) simulations in which the galaxies fall into a massive ``live'' group or cluster on realistic orbits. One important aim is to derive a physically simple and accurate description of the stripping seen in the simulations that can be easily employed in the modelling of observed or cosmologically-simulated galaxies. An additional motivation for deriving such a model is to improve the treatment of ram pressure stripping in semi-analytic models of galaxy formation. At present, these models typically assume that the hot gaseous halos of galaxies are stripped at the instant they cross the virial radius of the group or cluster. Clearly, this is not a realistic assumption, especially in the case where the mass of the galaxy is not negligible compared to that of the group or cluster. Such semi-analytic models tend to predict group and cluster galaxies whose colours are too red compared to observations (e.g., Weinmann et al.\\ 2006; Baldry et al.\\ 2006). If the ram pressure stripping of the hot gaseous halos of cluster galaxies is not as (maximally) efficient as assumed by these models, the resulting galaxies would be bluer and perhaps in better agreement with observations. The present paper is structured as follows. In \\S 2, we present a discussion of our simulation setup and the results of convergence tests that demonstrate the robustness of our findings. In \\S 3, we first outline a simple analytic model for ram pressure stripping that is based on the original formulation of Gun \\& Gott (1972) but which is appropriate for spherically-symmetrical gas distributions (as opposed to disks). We then compare this model to a wide variety of simulations and demonstrate that it provides an excellent match to the mass loss seen in the simulations. Finally, in \\S 4, we summarise and discuss our findings. ", "conclusions": "Using a suite of carefully controlled 3D hydrodynamic simulations, we have investigated the ram pressure stripping of hot gas in the halos of galaxies as they fall into groups and clusters. We have proposed a physically simple analytic model that describes the stripping seen in the simulations remarkably well. This model is analogous to the original formulation of Gunn \\& Gott (1972), except that it is appropriate for the case of a spherical gas distribution (as opposed to a face-on disk) and takes into account that stripping is not instantaneous but occurs on approximately a sound crossing time. The only pieces of information that the model requires are the initial conditions of the orbiting galaxy (its gas and dark matter profiles), the density profile of the ICM and the orbit [the latter two are needed to calculate $P_{\\rm ram}(t)$]. The model contains two tunable coefficients that are of order unity. Fixing these coefficients to match the stripping in just one of our idealised uniform medium simulations (see \\S 3.2) leads to excellent agreement with all our other simulations. With the exception of cases where the mass of the galaxy is greater than about 10\\% of the mass of the group and its orbit is highly non-radial, the analytic model reproduces the mass loss in the simulations to $\\approx 10\\%$ accuracy at all times and for all the orbits, galaxy masses, and galaxy concentrations that we have explored. For cases where the mass of the galaxy exceeds 10\\% of the mass of the group, it will likely be necessary to factor in the effects of tidal stripping and gravitational shock heating, which are neglected by our model. We re-iterate that the numerical simulations with which our analytic model has been calibrated have been demonstrated to be robust to the adopted resolution and artificial viscosity strength (see \\S 2.2). Furthermore, as we have demonstrated that KH (and RT) instability stripping is expected to be unimportant, SPH codes should be fully capable of tackling the problem of hot halo gas stripping in galaxies orbiting in groups and clusters. A direct comparison between the results using the GADGET-2 and FLASH hydrodynamic codes for one of our runs (see Appendix A) confirms this conclusion. The model we have derived has a number of potentially interesting applications, including modelling observed satellite galaxies and satellite galaxies in cosmological simulations. One application that we are currently pursuing is the incorporation of our ram pressure stripping model into a semi-analytic model of galaxy formation. As mentioned in \\S 1, recent observations (Weinmann et al.\\ 2006; Baldry et al.\\ 2006) have revealed that current semi-analytic models predict satellite galaxies whose colours are too red compared to the observed systems. The implementation of ram pressure stripping in these models is unrealistically efficient since, by assumption, the hot halo of the satellite galaxy is instantly transferred to the more massive system as soon as the satellite galaxy enters the massive system's virial radius. In reality, the hot gaseous halo of the galaxy will remain intact for a while. For example, for the most common orbital parameters, we find that between 20\\%-40\\% of the initial hot halo of the galaxy can remain in place even after 10 Gyr of orbiting inside the group or cluster (see Fig.\\ 9; note, however, that the quoted numbers could be sensitive to the adopted hot gas distribution of the galaxy). We note that these predictions are in qualitative agreement with recent {\\it Chandra} X-ray observations of massive galaxies orbiting in hot clusters by Sun et al.\\ (2007b), who find that most of the galaxies have detectable hot gaseous halos. Depending on the efficiency of feedback (e.g., from supernovae winds) in the semi-analytic models, radiative cooling of the remaining hot halo gas will replenish the cold gaseous component at the centre of the galaxy, which in turn will allow star formation to continue for some time. This will have the effect of making the colour of model satellite galaxies bluer and could resolve the discrepancy between semi-analytic models and observations (Font et al., in prep)." }, "0710/0710.1022.txt": { "abstract": "{ We deal with the test of the general relativistic gravitomagnetic Lense-Thirring effect currently being conducted in the Earth's gravitational field with the combined nodes $\\Omega$ of the laser-ranged geodetic satellites LAGEOS and LAGEOS II. One of the most important sources of systematic uncertainty on the orbits of the LAGEOS satellites, with respect to the Lense-Thirring signature, is the bias due to the even zonal harmonic coefficients $J_{\\ell}$ of the multipolar expansion of the Earth's geopotential which account for the departures from sphericity of the terrestrial gravitational potential induced by the centrifugal effects of its diurnal rotation. The issue addressed here is: are the so far published evaluations of such a systematic error reliable and realistic? The answer is negative. Indeed, if the difference $\\Delta J_{\\ell}$ among the even zonals estimated in different global solutions (EIGEN-GRACE02S, EIGEN-CG03C, GGM02S, GGM03S, ITG-Grace02, ITG-Grace03s, JEM01-RL03B, EGM2008, AIUB-GRACE01S) is assumed for the uncertainties $\\delta J_{\\ell}$ instead of using their more-or-less calibrated covariances $\\sigma_{J_\\ell}$, it turns out that the systematic error $\\delta\\mu$ in the Lense-Thirring measurement is about 3 to 4 times larger than in the evaluations so far published based on the use of the covariances of one model at a time separately, amounting up to $37\\%$ for the pair EIGEN-GRACE02S/ITG-Grace03s. The comparison among the other recent GRACE-based models yields bias as large as about $25-30\\%$. The major discrepancies still occur for $J_4, J_6$ and $J_8$, which are just to which the zonals the combined LAGEOS/LAGOES II nodes are most sensitive. } %% Keywords should be separated by \\*\\ sign ", "introduction": "In the weak-field and slow motion approximation, the Einstein field equations of general relativity get linearized to a form resembling Maxwell's equations of electromagnetism. Thus, a gravitomagnetic field, induced by the off-diagonal components $g_{0i}, i=1,2,3$ of the space-time metric tensor related to the mass-energy currents of the source of the gravitational field, arises \\cite{MashNOVA}. It affects in several ways the motion of, e.g., test particles and electromagnetic waves \\cite{Rug}. Perhaps the most famous gravitomagnetic effects are gyroscope precession \\cite{Pugh,Schi} and the Lense-Thirring\\footnote{According to an interesting historical analysis recently performed in Ref. \\cite{Pfi07}, it would be more correct to speak about an Einstein-Thirring-Lense effect.} precessions \\cite{LT} of the orbit of a test particle, both occurring in the field of a central slowly rotating mass like a planet. The measurement of gyroscope precession in the Earth's gravitational field has been the goal of the dedicated space-based GP-B mission\\footnote{See \\url{http://einstein.stanford.edu/}} \\cite{Eve, GPB} launched in 2004; its data analysis is still in progress. In this paper we critically discuss some issues concerning the test of the Lense-Thirring effect performed with the LAGEOS and LAGEOS II terrestrial artificial satellites \\cite{Ciu04} tracked with the Satellite Laser Ranging (SLR) technique \\cite{Pearl02}. \\cite{vpe76a,vpe76b} proposed measuring the Lense-Thirring nodal precession of a pair of counter-orbiting spacecraft in terrestrial polar orbits and equipped with drag-free apparatus. A somewhat equivalent, cheaper version of such an idea was put forth in Ref. \\cite{Ciu86} whose author proposed to launch a passive, geodetic satellite in an orbit identical to that of the LAGEOS satellite apart from the orbital planes which should have been displaced by 180 deg apart.\\footnote{ LAGEOS was put into orbit in 1976, followed by its twin LAGEOS II in 1992.} The measurable quantity was, in this case, the sum of the nodes of LAGEOS and of the new spacecraft, later named LAGEOS III, LARES, WEBER-SAT, in order to cancel the confounding effects of the multipoles of the Newtonian part of the terrestrial gravitational potential (see below). Although extensively studied by various groups \\cite{CSR,LARES,Ioretal02}, such an idea has not been implemented for a long time. Recently, the Italian Space Agency (ASI) has approved this project and should launch a VEGA rocket for this purpose at the end of 2009-beginning of 2010 (\\url{http://www.asi.it/en/activity/cosmology/lares}). For recent updates of the LARES/WEBER-SAT mission, including recently added additional goals in fundamental physics and related criticisms, see Refs. \\cite{LucPao01,Ior02,Ciu04b,Ciu06b,IorJCAP,Ior07c,Ior07d,Ior09}. Among scenarios involving \\emph{existing} orbiting bodies, the idea of measuring the Lense-Thirring node rate with the just launched LAGEOS satellite, along with the other SLR targets orbiting at that time, was proposed in Ref. \\cite{Cug78}. Tests have been effectively performed using the LAGEOS and LAGEOS II satellites \\cite{tanti}, according to a strategy \\cite{Ciu96} involving a suitable combination of the nodes of both satellites and the perigee $\\omega$ of LAGEOS II. This was done to reduce the impact of the most relevant source of systematic bias, viz. the mismodelling in the even ($\\ell=2,4,6\\ldots$) zonal ($m=0$) harmonics $J_{\\ell}$ of the multipolar expansion of the Newtonian part of the terrestrial gravitational potential:\\footnote{The relation among the even zonals $J_{\\ell}$ and the normalized gravity coefficients $\\overline{C}_{\\ell 0}$ is $J_{\\ell}=-\\sqrt{2\\ell + 1}\\ \\overline{C}_{\\ell 0}$.} they account for non-sphericity of the terrestrial gravitational field induced by centrifugal effects of the Earth's diurnal rotation. The even zonals affect the node and the perigee of a terrestrial satellite with secular precessions which may mimic the Lense-Thirring signature. The three-elements combination used allowed for removing the uncertainties in $J_2$ and $J_4$. In \\cite{Ciu98} a $\\approx 20\\%$ test was reported by using the\\footnote{Contrary to the subsequent CHAMP/GRACE-based models, EGM96 relies upon multidecadal tracking of SLR data of a constellation of geodetic satellites including LAGEOS and LAGEOS II as well; thus the possibility of a sort of $a-priori$ `imprinting' of the Lense-Thirring effect itself, not solved-for in EGM96, cannot be neglected.} EGM96 \\cite{Lem98} Earth gravity model; subsequent analyses showed that such an evaluation of the total error budget was overly optimistic in view of the likely unreliable computation of the total bias due to the even zonals \\cite{Ior03,Ries03a,Ries03b}. An analogous, huge underestimation turned out to hold also for the effect of non-gravitational perturbations \\cite{Mil87} like direct solar radiation pressure, the Earth's albedo, various subtle thermal effects depending on the the physical properties of the satellites' surfaces and their rotational state \\cite{Inv94,Ves99,Luc01,Luc02,Luc03,Luc04,Lucetal04,Ries03a}, which the perigees of LAGEOS-like satellites are particularly sensitive to. As a result, the realistic total error budget in the test reported in Ref. \\cite{Ciu98} might be as large as $60-90\\%$ or even more (by considering EGM96 only). The observable used in Ref. \\cite{Ciu04} with the GRACE-only EIGEN-GRACE02S model \\cite{eigengrace02s} and in Ref. \\cite{Ries08} with other global terrestrial gravity solutions was the following linear combination\\footnote{See also \\cite{Pav02,Ries03a,Ries03b}.} of the nodes of LAGEOS and LAGEOS II, explicitly computed in Ref. \\cite{IorMor} following the approach proposed in Ref. \\cite{Ciu96}: \\begin{equation} f=\\dot\\Omega^{\\rm LAGEOS}+ c_1\\dot\\Omega^{\\rm LAGEOS\\ II }, \\lb{combi}\\end{equation} where \\begin{equation} c_1\\equiv-\\rp{\\dot\\Omega^{\\rm LAGEOS}_{.2}}{\\dot\\Omega^{\\rm LAGEOS\\ II }_{.2}}=-\\rp{\\cos i_{\\rm LAGEOS}}{\\cos i_{\\rm LAGEOS\\ II}}\\left(\\rp{1-e^2_{\\rm LAGEOS\\ II}}{1-e^2_{\\rm LAGEOS}}\\right)^2\\left(\\rp{a_{\\rm LAGEOS\\ II}}{a_{\\rm LAGEOS}}\\right)^{7/2}.\\lb{coff}\\end{equation} The coefficients $\\dot\\Omega_{.\\ell}$ of the aliasing classical node precessions \\cite{Kau} $\\dot\\Omega_{\\rm class}=\\sum_{\\ell}\\dot\\Omega_{.\\ell}J_{\\ell}$ induced by even zonals have been analytically worked out in e.g. \\cite{Ior03}; $a,e,i$ are the satellite's semimajor axis, eccentricity and inclination, respectively and yield $c_1=0.544$ for \\rfr{coff}. The Lense-Thirring signature of \\rfr{combi} amounts to 47.8 milliarcseconds per year (mas yr$^{-1}$). The combination \\rfr{combi} allows, by construction, to remove the aliasing effects due to the static and time-varying parts of the first even zonal $J_2$. The nominal bias (computed with the estimated values of $J_{\\ell}$, $\\ell=4,6...$) due to the remaining higher degree even zonals would amount to about $10^5$ mas~yr$^{-1}$; the need of a careful and reliable modeling of such an important source of systematic bias is, thus, quite apparent. Conversely, the nodes of the LAGEOS-type spacecraft are affected by the non-gravitational accelerations $\\approx 1\\%$ of the Lense-Thirring effect \\cite{Luc01,Luc02,Luc03,Luc04,Lucetal04}. For a comprehensive, up-to-date overview of the numerous and subtle issues concerning the measurement of the Lense-Thirring effect see \\cite{IorNOVA}. Here, we will address the following questions: \\begin{itemize} \\item Has the systematic error due to the competing secular node precessions induced by the static part of the even zonal harmonics been realistically evaluated so far in literature? (Section \\ref{grav}) % \\item Why has the analysis with the LAGEOS satellites not been % repeated so far by any other independent team? (\\sref{repeat}) \\item Are other approaches to extract the gravitomagnetic signature from the data feasible? (Section \\ref{approach}) \\end{itemize} ", "conclusions": "In this paper we have shown how the so far published evaluations of the total systematic error in the Lense-Thirring measurement with the combined nodes of the LAGEOS satellites due to the classical node precessions induced by the even zonal harmonics of the geopotential are likely optimistic. Indeed, they are all based on the use of elements from the covariance matrix, more or less reliably calibrated, of various Earth gravity model solutions used one at a time separately in such a way that the model X yields an error of $x\\%$, the model Y yields an error $y\\%$, etc. Instead, comparing the estimated values of the even zonals for different pairs of models allows for a much more realistic evaluation of the real uncertainties in our knowledge of the static part of the geopotential. As a consequence, the bias in the Lense-Thirring effect measurement is about three to four times larger than that so far claimed, amounting to tens of parts per cent (37$\\%$ for the pair EIGEN-GRACE02S and ITG-GRACE03s, about 25--30$\\%$ for the other most recent GRACE-based solutions). % % We have also pointed out that, until now, no other tests of the % Lense-Thirring effect have been performed by independent teams, % although it would % be, %at least in principle, a relatively not too demanding task in view of the wide dissemination of SLR stations, for which the LAGEOS satellites are %important targets since long time, and of the freely available softwares to perform the data reduction process. % Finally, we have pointed out the need of following different strategies in extracting the Lense-Thirring pattern from the data; for instance by explicitly modelling it in fitting the SLR data of LAGEOS and LAGEOS II, and estimating the associated solve-for parameter in a least-square sense along with the other parameters usually determined. %" }, "0710/0710.0141_arXiv.txt": { "abstract": "{} {Molecules that trace the high-density regions of the interstellar medium have been observed in (ultra-)luminous (far-)infrared galaxies, in order to initiate multiple-molecule multiple-transition studies to evaluate the physical and chemical environment of the nuclear medium and its response to the ongoing nuclear activity.} {The HCN(1$-$0), HNC(1$-$0), \\HCOP(1$-$0), CN(1$-$0) and CN(2$-$1), CO(2$-$1), and CS(3$-$2) transitions were observed in sources covering three decades of infrared luminosity including sources with known OH megamaser activity. The data for the molecules that trace the high-density regions have been augmented with data available in the literature.} {The integrated emissions of high-density tracer molecules show a strong relation to the far-infrared luminosity. Ratios of integrated line luminosities have been used for a first order diagnosis of the integrated molecular environment of the evolving nuclear starbursts. Diagnostic diagrams display significant differentiation among the sources that relate to initial conditions and the radiative excitation environment. Initial differentiation has been introduced between the FUV radiation field in photon-dominated-regions and the X-ray field in X-ray-dominated-regions. The galaxies displaying OH megamaser activity have line ratios typical of photon-dominated regions. } {} ", "introduction": "Large amounts of high-density molecular gas play a crucial role in the physics of (ultra-)luminous (far-)infrared galaxies ((U)LIRGs), giving rise to spectacular starbursts and possibly providing the fuel for the emergence of active galactic nuclei (AGN). A strong linear relation has been found between the far infrared (\\irlums) and HCN luminosity, the latter being an indicator of the high density (\\numd$\\ga$10$^4$\\,\\percc) component of the interstellar medium. This arises predominantly from the nuclear region and indicates a close relation to the nuclear starburst environment and the production of the FIR luminosity \\citep{GaoS2004a}. The molecular gas contributes a significant fraction to the dynamical mass of the central regions of FIR galaxies \\citep{1991A&ARv...3...47H, YoungScoville1991}. Emissions of molecules that trace the high-density regions in LIRGs and ULIRGs may serve to study the characteristics of the extreme environments in the nuclear regions of starburst and AGN nuclei. While much of the energy generation of FIR-luminous galaxies has traditionally been attributed to AGN activity using their optical signatures, the NIR and radio signatures suggest significant starburst contributions \\citep{GenzelEA1998, BaanKlockner2006}. Circum-nuclear star formation in the densest regions serves as a power source for the majority of ULIRGs such that they even mimic the presence of an AGN \\citep{BaanKlockner2006}. The ULIRG population is also characterized by OH megamasers (MM) \\citep{Baan1989, HenkelWilson1990} resulting from amplification of radio continuum by FIR-pumped foreground gas. Few H$_2$CO MM are also found among this population \\citep{BaanHU1993, ArayaBH2004}. The observed molecular emissions from the high-density components could carry the signature of the nuclear region of the ULIRG and its nuclear power generation. Recent studies consider the HCN\\,(1$-$0) versus CO\\,(1$-$0) relation and its implications for the star formation activity \\citep{GaoS2004a, GaoS2004b}. Discussions are ongoing in the literature about whether such molecular emissions are accurate tracers of the high-density medium at the site of star formation \\citep{WuEvansEA2005, GraciaCarpioGPC2006, Papadop2007}. It should be evident that only multi-transition and multi-molecule (multi-dimensional) studies can describe the multi-parameter environment of the nuclear activity and further establish the connection between the molecular signature and the nature of the nuclear energy source. The translation or extrapolation from the characteristics of individual star formation regions in our Galaxy to integrated regions in nearby galaxies at larger distances or even to unresolved nuclear regions in distant galaxies requires detailed multi-species observations and comparisons combined with elaborate theoretical modelling. The interpretation of molecular line data has focused on the nature of the local excitation mechanisms in the form of photon dominated regions (PDR) with far-ultraviolet radiation fields and X-ray dominated regions (XDR) \\citep{1996A&A...306L..21L, MeijerinkS2005}. The molecular properties of these two different regimes can be used as diagnostic tools in interpreting the integrated properties of galaxies. This paper reflects a multi-dimensional approach to understanding molecular line emissions that started in the early 90's, but is only now coming to fruition. This study originally aimed to establish a better link between molecular megamaser activity and the molecular properties of luminous FIR galaxies but it has changed into a general study of ULIRGs. Here we present data on the molecular characteristics of normal and FIR galaxies across the available range of \\irlums. Data of the CO, HCN, HNC, \\HCOP, CN, and CS line emissions obtained for a representative group of 37 FIR-luminous and OH megamaser galaxies has been joined with the additional data of 80 sources presented in the literature. Early studies of the properties of CS \\citep{MauersbergerHWH1989}, \\HCOP\\ \\citep{NguyenEA1992}, and HNC \\citep{HuttemeisterEA1995} using small numbers of sources studied the relations between line and FIR luminosities and the nature of the central engine. Our molecular emission data show clear tendencies that cover a wide regime of nuclear activity. This database forms the basis for a first evaluation of the emission line properties and for further study and modelling of the line properties. ", "conclusions": "Molecular line emissions in FIR-luminous galaxies are tools for multi-dimensional diagnostics of the environmental parameters in the nuclear ISM and the heating processes resulting from the nuclear activity. A total of 118 line detections made with the 30-m Pico Veleta and the 15-m SEST telescopes for a total of 37 sources have been complemented with partial records published for another 80 sources. A proper evaluation of integrated emission lines and their diagnostic line ratios can be achieved after synthesis and modelling of a representative description of the integrated nuclear molecular medium under a wide variety of AGN- and starburst-related circumstances. The molecular information in this paper has been diagnosed to first order using modelling results for nuclear emission regions presented in the literature. The collective behavior of line luminosities and line ratios of the low-density and high-density tracers presents a consistent picture of the molecular medium in the nuclear regions of ULIRGs. The high-density tracers represent the molecular medium in the regions where star formation is taking place, and the low-density tracer represents the relatively unperturbed larger-scale molecular environment. The tracer luminosities increase roughly linearly with FIR luminosity but they show significant scatter in data points due to physical processes in the nuclear region, the chemical history, and the relative age of the nuclear activity. The luminosity of high-density tracers varies linearly with \\irlum except for the CS(3$-$2) luminosity, where a steeper relation is suggestive of a different excitation dependence. The CO(1$-$0) and CO(2$-$1) lines for these high FIR luminosity sources have a slope less than unity and are (almost) consistent with empirical evidence on the relation of SFR versus gas content \\citep{Kennicutt1998}. First order diagnostics of the nuclear ISM can be based on the collective behavior of high-density tracers HCN, HNC and \\HCOP\\ showing significant differentiation and systematic changes. The line strengths of these three molecules relative to the CO\\,(1$-$0) line show a significant range of 0.03 to 0.23 with a distinct dependence on \\irlums. This dependence has been interpreted in terms of an evolutionary model where the depletion of molecular gas depends on the consumption and destruction of the high-density gas by the ongoing star formation process. Only partial diagnostics could be done using the available CN and CS data, which complements the diagnostics of the three other high-density tracers. The emission line ratios of the three high-density tracers of the galaxies show a structured distribution that fills selected parts of the parameter space. Interpreting these distributions has been done using modelling column densities that are characteristic of PDR- and XDR-dominated nuclear environments. OH MM and other powerful ULIRGs are mostly characterized by PDR-dominance. The other end of the data distribution is characterized by mixed PDR-XDR and XDR-dominated environments. (U)LIRGs represent the phases of nuclear evolution that generate the highest FIR luminosities and most rapidly deplete the dense molecular component. Using the HCN/CO ratio as an evolutionary indicator, the distributions of data points may reflect evolution from PDR-like to more XDR-like nuclear ISM properties during the course of the outburst. The \\HCOP/HCN and HNC/HCN ratios serve as indicators of environments with shocks and non-standard heating and with HNC depletion and \\HCOP\\ enhancement in a fraction of the sources. The \\HCOP/HCN ratio also serves as a density indicator and would be higher under PDR circumstances and lower in XDR environments. The observed trends in the \\HCOP/HCN ratio could result from the (indirect) influence of an AGN in the nucleus. More likely the trends in the relative abundance of HCN as compared with other constituents result from evolution of the nuclear environment during the SBN activity. The detailed interpretations of multi-line multi-molecule emission line behavior and their relation with the local environment require detailed modelling of the physical parameters of the environment together with the excitation, the chemistry, and the radiative transfer for the molecular constituents. Hereby the integration of higher level transitions of the molecules is needed to determine specific excitation temperatures and densities. A comparison with the properties of Galactic emission regions will emphasize the effect of scale size on the line ratios for tracer molecules. Further studies are underway to connect the FIR signature and global heating scenario of ULIRGs \\citep{LoenenBS2006} with modelling of the molecular emissions under varying conditions in a nuclear starburst. The emission scenarios for XDRs and PDRs establish the connection between star formation and other sources of excitation and the integrated emission line parameters observed in extragalactic sources. \\bigskip" }, "0710/0710.5849_arXiv.txt": { "abstract": "{ The Edelweiss programme is dedicated to the direct search for Dark Matter as massive weakly interacting particles (WIMPs) with Germanium cryogenic detectors operated in the Laboratoire Souterrain de Modane in the French Alps at a depth of 4800 mwe. After the initial phase Edelweiss I, which involved a total mass of 1 kg, the second step of the programme, Edelweiss II, currently operates 9 kg of detectors and an active shielding of 100 m$^2$ muon veto detectors and is now in its commissioning phase. The current status and performance of the Edelweiss II set-up in terms of backgrounds will be given, the underground muon flux measured with the muon veto system will be presented. \\PACS{ {95.35.+d}{Dark matter search} \\and {14.80.Ly}{Supersymmetric partners of known particles} \\and {98.80.Es}{Observational cosmology} \\and {98.70.Vc}{Background radiations, cosmic} } % } % ", "introduction": "\\label{intro} Understanding the nature of Dark Matter in the universe is a major challenge for modern cosmology and astrophysics. One of the well-motivated candidates is the generically named Weakly Interacting Massive Particle (WIMP). In the Minimal Supersymetric Standard Model (MSSM) framework, the WIMP could be the Lightest Supersymmetric Particle, which is stable, neutral and massive. The EDELWEISS (Experience pour DEtecter Les Wimps En Site Souterrain) experiment is dedicated to the direct detection of WIMPs trapped in the galactic halo. The detection is performed through the measurement of the recoil energy produced by elastic scattering of a WIMP off target nuclei. The main challenges are the extremely low event rate of $\\le$ 1 evt/kg/year and the relatively small deposited energy of $\\le$ 100 keV. ", "conclusions": "\\label{sec:conc} The EDELWEISS-II setup has been validated with calibration and background runs. Energy resolutions and discrimination capabilities close to those of EDEL\\-WEISS-I have been measured for Ge/NTD detectors. Validation of Ge/NbSi detectors with new aluminium electrodes is in progress and Ge/NTD detectors with interdigitized electrodes scheme have shown promising results in surface laboratory. The muon veto system runs, and study of coincidences between the muon veto and the bolometer systems will be performed in the following months. A new setup for muon-neutron measurement will be soon installed at LSM and will allow to evaluate the muon induced neutron flux. Low background physics runs will be taken with the 28 bolometers setup with the aim to reach sensitivity to WIMP-nucleon cross-section of $\\sim$10$^{-7}$ pb for a WIMP mass of 100 GeV by July 2008. 40 additional detectors (Ge/NbSi and Ge/NTD with interdigitized electrodes) will be added in the two coming years to enhance progressively the sensitivity to $\\sim$10$^{-8}$ pb thanks to active surface rejection, with an expected goal reached by autumn 2009/2010. \\sloppy \\begin{acknowledgement} {\\bf Acknowledgment:} This work has been partially supported by the EEC Applied Cryodetector network (Contract HPRN-CT-2002-00322), the ILIAS integrating activity (Contract RII3-CT-2003-506222), the HGF initiative and networking fund through the virtual institute VIDMAN and the SFB/Transregio 27 of the Deutsche Forschungsgemeinschaft, DFG. The help of the technical staff of the Laboratoire Souterrain de Modane and of the participating laboratories is gratefully acknowledged. \\end{acknowledgement}" }, "0710/0710.2691_arXiv.txt": { "abstract": "We present new optical, near-IR, and mid-IR observations of the young eruptive variable star V1647~Orionis that went into outburst in late 2004 for approximately two years. Our observations were taken one year after the star had faded to its pre-outburst optical brightness and show that V1647~Ori is still actively accreting circumstellar material. We compare and contrast these data with existing observations of the source from both pre-outburst and outburst phases. From near-IR spectroscopy we identify photospheric absorption features for the first time that allow us to constrain the classification of the young star itself. Our best fit spectral type is M0$\\pm$2 sub-classes with a visual extinction of 19$\\pm$2 magnitudes and a K-band veiling of r$_K\\sim$1.5$\\pm$0.2. We estimate that V1647~Ori has a quiescent bolometric luminosity of $\\sim$9.5~L$_{\\odot}$ and a mass accretion rate of $\\sim$1$\\times$10$^{-6}$~M$_\\odot$~yr$^{-1}$. Our derived mass and age, from comparison with evolutionary models, are 0.8$\\pm$0.2~M$_{\\odot}$ and $\\lesssim$~0.5~Myrs, respectively. The presence towards the star of shock excited optical [S~II] and [Fe~II] emission as well as near-IR H$_2$ and [Fe~II] emission perhaps suggests that a new Herbig-Haro flow is becoming visible close to the star. ", "introduction": "A significant event in the recent history of star formation studies occurred in late 2003 when the young pre-main sequence star V1647~Orionis went into outburst. This eruption, and the associated one hundred fold increase in optical brightness, resulted in the appearance of a monopolar reflection nebula now known as ''McNeil's Nebula'' after the amateur astronomer, Jay McNeil, who discovered it (McNeil 2004). The star and nebula remained bright for approximately 18 months before fading rapidly over a six month period. V1647~Ori was found to be again close to its pre-outburst (Sloan Digital Sky Survey, SDSS) optical brightness in early 2006. A widely accepted interpretation of such eruptions is one involving a heated disk and a `mass dumping' episode from a circumstellar disk onto a stellar photosphere. The five magnitude increase in optical brightness seen in V1647~Ori has been attributed to both the addition of significant accretion luminosity and a partial dust clearing due to the high-velocity wind (up to 600~km~s$^{-1}$) emanating from the star/disk engine (Reipurth \\& Aspin 2004; McGehee et al. 2004). Both star and nebula have been widely studied from the period soon after the outburst occurred through to its fading to its pre-outburst brightness. In our discussions below, we refer to the period before November 2004 as the {\\it pre-outburst phase} (Brice\\~no et al. 2004), the period from November 2004 to February 2006 as the {\\it outburst phase}, and the period from February 2006 onwards as the {\\it quiescent phase} of V1647~Ori. Observations spanning all spectral regimes, from x-ray into the optical to the near-IR and on to the far-IR/sub-mm/radio, have been published by many authors in a large number of papers. Readers are referred to these for detailed background information on V1647~Ori and McNeil's Nebula. These papers are, in alphabetical order: \\'Abr\\'aham et al. (2004 - the pre-outburst infrared characteristics), Acosta-Pulido et al. (2007 - optical and near-IR imaging photometry, near-IR spectroscopy and polarimetry during the outburst, henceforth AP07), Andrews, Rothberg, \\& Simon (2004 -- mid-IR and sub-mm observations during the outburst), Aspin et al. (2006 -- an historical study of previous outbursts), Brice\\~no et al. (2004 -- the optical outburst history, pre-outburst and during the outburst), Fedele et al. (2007a,b -- optical and near-IR properties during the outburst), Gibb et al. (2006 -- near-IR spectroscopy during the outburst), Grosso et al. (2005 -- x-ray observations during the outburst), Kastner et al. (2004 -- x-ray emission pre-outburst and during the outburst), Kastner et al. (2006 -- x-ray evolution during the outburst), K\\'osp\\'al et al. (2005 -- optical photometry during the outburst), McGehee et al. (2004 -- optical and near-IR photometry pre-outburst and during the outburst), Mosoni et al. (2005 -- mid-IR interferometric observations during the outburst), Muzerolle et al. (2005 -- {\\it Spitzer} observations during the outburst), Ojha et al. (2004 -- near-IR imaging during the outburst), Ojha et al. (2006 -- optical photometry and spectroscopy, and near-IR imaging during the outburst), Reipurth \\& Aspin (2004 -- optical and near-IR imaging photometry and spectroscopy during the outburst), Rettig et al. (2005 -- high spectral resolution near-IR observations during the outburst), Semkov (2004 -- optical photometry during the outburst), Semkov (2006 -- optical light curve during the outburst), Tsukagoshi et al. (2005 -- mm continuum observations during the outburst), Vacca, Cushing, \\& Simon (2004 -- near-IR spectroscopy during the outburst), Vig et al. (2006 -- radio observations during the outburst), and Walter et al. (2004 -- optical photometry and spectroscopy during the outburst). The significance of eruptive outbursts, similar to the one undergone by V1647~Ori lies in the fact that these periods are considered times when mass accretion rates increase dramatically (Hartmann \\& Kenyon 1996, henceforth HK96). During periods of outburst, mass accretion rates have been directly observed to increase by up to several orders of magnitude from a pseudo-steady-state value of $\\sim$10$^{-9~to~-7}$~M$_\\odot$~yr$^{-1}$ for classical T~Tauri stars (Hartmann et al. 1998) to $\\sim$10$^{-6~to~-5}$~M$_\\odot$~yr$^{-1}$ for EXor variables, and as high as $\\sim$10$^{-4}$~M$_\\odot$~yr$^{-1}$ for the extremely energetic FUor variables (Hartmann \\& Kenyon 1985). FUors and EXors are so named after the prototypes of their classes, namely FU~Orionis and EX~Lupi, respectively (Herbig 1989). Whether all young stars undergo EXor and/or FUor eruptions and, if so, whether these two types of outbursts events are fundamentally the same (on perhaps different scales) or different (in terms of the mechanism involved and the triggering of the eruption) is still a matter of debate. EXor eruptions are distinct from FUor eruptions in that they are much shorter lived, months to years rather than decades to centuries, and have been empirically determined to be repetitive (Herbig 1989; 2007). For example, EX~Lupi itself has gone into outburst on several occasions, with the last being in 1993--94 (Lehmann et al. 1995, Herbig et al. 2001). V1647~Ori is likely another example of an EXor since its recent activity lasted only two years and it was previously observed in outburst from 1966--67 (Aspin et al. 2006) on archival photographic plates. Mechanisms that have been proposed to initiate an eruptive event, be it EXor-like or FUor-like, include: thermal disk instabilities resulting in a runaway accretion condition (Bell \\& Lin 1994), periodic overloading of the inner regions of the circumstellar disk and subsequent magnetic collapse (HK96), and the close approach of a companion in an eccentric orbit disturbing the stability of the inner regions of the disk (Bonnell \\& Bastien 1992). In this paper, we present new optical, near-IR, and mid-IR observations of V1647~Ori and McNeil's Nebula taken approximately one year after the source reached its pre-outburst optical brightness. Below, we compare and contrast our observations with those taken during the pre-outburst and outburst phases of the eruption and attempt to better understand the nature and characteristics of the underlying young star and its circumstellar environment. ", "conclusions": "From our optical, near-IR, and mid-IR imaging and spectroscopy of V1647~Ori dating from February to April 2007, approximately one year after its return to its pre-outburst optical brightness, we can conclude that: \\begin{itemize} \\item The associated nebula, McNeil's Nebula, remained faintly visible suggesting that quasi-static nebula material is still being illuminated by V1647~Ori. \\item We confirm the findings of Fedele et al. (2007a) in that signposts of shock-excited emission, specifically, in the emission lines of [S~II], [Fe~II], and H$_2$ are present. We consider that it is a distinct possibility that these emission lines result from a new Herbig-Haro flow. \\item Our near-IR spectrum shows, for the first time, molecular overtone absorption from CO and atomic absorption from neutral Na and Ca. We interpret this as evidence that we are now observing the stellar photosphere of V1647~Ori. \\item We have modeled the near-IR spectrum using a template stellar spectrum including both water vapor and water ice absorption and have determined a best-fit parameter set of spectral type M0$\\pm$2 sub-classes, A$_V$=19$\\pm$2 magnitudes, r$_K$=1.5$\\pm$0.2. \\item We derive values of M$_{bol}$=2.9$\\pm$0.4 magnitudes and L$_{*}$=5.2$\\pm$2~L$_{\\odot}$ for V1647~Ori. From comparison with theoretical evolutionary tracks, find that, for the adopted T$_{eff}\\sim$3800~K, the star has a mass and age of 0.8$\\pm$0.2~M$_{\\odot}$ and $\\lesssim$0.5~Myrs, respectively. \\item From near-IR H~I line flux and the A$_V$ values derived above, we estimate the accretion luminosity and mass accretion rate in February 2007 was L$_{acc}$=4.0$\\pm$2~L$_{\\odot}$ and \\.M=1.0$\\times$10$^{-6}$~M$_{\\odot}$~yr$^{-1}$, respectively. This implies that V1647~Ori is still actively accreting circumstellar material even though it is almost optically invisible and that the accretion rate during the outburst must have been considerably larger that this value. \\item V1647~Ori is found to have a quiescent phase L$_{bol}$ of 9.25$\\pm$3~L$_{\\odot}$. \\item For the first time, we see evidence for silicate dust evolution in the mid-IR spectrum of V1647~Ori over the outburst to quiescence period. We now observe weak silicate absorption at 10~$\\mu$m whereas previously the silicate band was either absent or weakly in emission. \\item Finally we note that, in February 2007, the spectral energy distribution, SED, of V1647~Ori appears remarkably similar to its pre-outburst SED suggesting that perhaps the derived accretion rate is the normal quiescent phase accretion rate for this object. \\end{itemize} \\vspace{0.3cm} {\\bf Acknowledgments} Based on observations obtained at the Gemini Observatory (under program identification GN-2007A-Q-33 and GS-2005B-Q-13), which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the Particle Physics and Astronomy Research Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), CNPq (Brazil) and CONICET (Argentina). Colin Aspin and Tracy Beck were visiting astronomers at the Infrared Telescope Facility, which is operated by the University of Hawaii under Cooperative Agreement no. NCC 5-538 with the National Aeronautics and Space Administration, Science Mission Directorate, Planetary Astronomy Program. This material is based upon work supported by the National Aeronautics and Space Administration through the NASA Astrobiology Institute under Cooperative Agreement No. NNA04CC08A issued through the Office of Space Science." }, "0710/0710.0694_arXiv.txt": { "abstract": "We have designed a system to stabilize the gain of a submillimeter heterodyne receiver against thermal fluctuations of the mixing element. In the most sensitive heterodyne receivers, the mixer is usually cooled to 4 K using a closed-cycle cryocooler, which can introduce $\\sim$1\\% fluctuations in the physical temperature of the receiver components. We compensate for the resulting mixer conversion gain fluctuations by monitoring the physical temperature of the mixer and adjusting the gain of the intermediate frequency (IF) amplifier that immediately follows the mixer. This IF power stabilization scheme, developed for use at the Submillimeter Array (SMA), a submillimeter interferometer telescope on Mauna Kea in Hawaii, routinely achieves a receiver gain stability of 1 part in 6,000 (rms to mean). This is an order of magnitude improvement over the typical uncorrected stability of 1 part in a few hundred. Our gain stabilization scheme is a useful addition to SIS heterodyne receivers that are cooled using closed-cycle cryocoolers in which the 4 K temperature fluctuations tend to be the leading cause of IF power fluctuations. ", "introduction": "\\PARstart{S}{uperconductor} insulator superconductor (SIS) receivers are in use on many millimeter and submillimeter telescopes because of their good spectral line sensitivity. Their continuum sensitivity, however, does not usually reach the theoretical limit because of receiver gain fluctuations. These arise predominantly from changes in mixer conversion gain, which result from physical temperature changes and local oscillator (LO) power changes. Here we describe a technique to reduce the impact of physical temperature changes on mixer conversion gain variations by changing the intermediate frequency (IF) gain of the receiver in proportion to fluctuations of the physical temperature of the SIS junction used as a mixing element. A number of radio observatories make use of liquid Helium filled cryostats so that temperature induced receiver gain fluctuations are minimized. However, due to their lower degree of maintenance and upkeep, closed-cycle Helium cryocooler systems may be preferred. For example multi-receiver systems with high heat loads, such as those in use at the 8-element Submillimeter Array (SMA)\\footnote{The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics, and is funded by the Smithsonian Institution and the Academia Sinica.} \\cite{refs:SMA} or the upcoming 64-element Atacama Large Millimeter Array (ALMA) \\cite{refs:Wootten}, are dependent on the use of cryocoolers. For these applications it is necessary to develop techniques to reconcile the need for highly stable receivers with the practical benefits of cryocoolers. An obvious approach is to design more stable cryocoolers, another is to compensate for the fluctuations in the mixer's conversion gain. In the Berkeley-Illinois-Maryland-Association Millimeter Array (BIMA) a heating resistor on the cold stage is used to compensate for cryocooler temperature fluctuations \\cite{refs:plambeck}. This reduces long time scale temperature drifts, giving 1 mK rms temperature stability over several hours when the data are averaged in 2 second blocks. The cycling of the cryocooler's displacer also introduces temperature fluctuations on 1-2 second time scales, which are not treated by this technique. To provide a large thermal capacity that will damp the fluctuations in the temperature of the cryocooler, it is possible to add a liquid helium reservoir. Using this technique, a ten fold improvement in temperature stability, from 200 mK to 20 mK (peak to peak), has been demonstrated \\cite{refs:sekimoto}. The ALMA cryocooler also incorporates a helium reservoir and has achieved an rms temperature stability of 1 mK at 4 K \\cite{refs:anna}. We take a different approach to receiver stabilization. While monitoring the physical temperature of the mixer we actively vary the IF gain of the receiver to compensate for mixer conversion gain variations. Such a scheme is straightforward to implement and can be applied to existing systems without redesigning the cryocooler. We have developed and tested hardware and accompanying software to adjust the third stage gain of the low noise HEMT amplifier immediately following the mixer in proportion to the fluctuations in the physical temperature of the mixer. We achieve nearly an order of magnitude improvement in receiver stability from 1 part in $\\sim$800 to 1 part in 6,000 (rms to mean ratio) using a 33 millisecond integration time over 10 minutes. This level of receiver stability equals that which would be obtained if rms fluctuations of the physical temperature of the mixer were held to $<$ 2 mK. ", "conclusions": "We have developed a system that stabilizes the gain of a 230 GHz SIS heterodyne receiver to 1 part in 6,000 (rms to mean). With the use of an open-loop proportional servo system, we monitor the physical temperature of the mixer block and adjust the gain of the third and final stage of a low noise HEMT IF amplifier in order to compensate for the subsequent variations in the conversion gain of the mixer. We have shown that the conversion gain of the mixer varies linearly over the range of typical cryocooler-induced physical temperature fluctuations, and that the gain of the HEMT amplifier varies linearly with its third stage bias current. Using our gain stabilization system, we routinely achieve a total receiver gain stability of 1 part in 6,000 (rms to mean ratio) for a 0.033 second integration time over 10 minutes, which corresponds to an effective rms temperature fluctuation of 1.7 mK. In comparison, the typical fluctuations in the physical temperature of the mixer at the SMA are on the order of 50-100 mK, and the corresponding power fluctuations are 1 part in several hundred. Therefore, our gain stabilization servo system can provide more than an order of magnitude improvement over a typical unstabilized system. Our technique introduces a negligible level of instrumental phase error into the source signal. This gain stability system can be implemented on any heterodyne receiver system in which the physical temperature of the mixer can be monitored and the IF amplification can be adjusted in real time." }, "0710/0710.0007_arXiv.txt": { "abstract": "The Hubble Ultra Deep Field (HUDF) contains a significant number of \\bb-, \\vv- and \\ii-band dropout objects, many of which were recently confirmed to be young star-forming galaxies at $z\\!\\simeq\\!4\\!-\\!6$. These galaxies are too faint individually to accurately measure their radial surface brightness profiles. Their average light profiles are potentially of great interest, since they may contain clues to the time since the onset of significant galaxy assembly. We separately co-add \\vv, \\ii- and \\zz-band HUDF images of sets of $z\\!\\simeq\\!4,5$ and $6$ objects, pre-selected to have nearly identical compact sizes and the roundest shapes. From these stacked images, we are able to study the average(d) radial structure of these objects at much higher signal-to-noise ratio than possible for an individual faint object. Here we explore the reliability and usefulness of a stacking technique of compact objects at $z\\!\\simeq\\!4\\!-\\!6$ in the HUDF. Our results are: (1) image stacking provides reliable and reproducible average surface brightness profiles; (2) the shape of the average surface brightness profile shows that even the faintest $z\\!\\simeq\\!4\\!-\\!6$ objects are \\emph{resolved}; and (3) if late-type galaxies dominate the population of galaxies at $z\\!\\simeq\\!4\\!-\\!6$, as previous \\emph{HST} studies have shown for $z\\!\\lesssim\\!4$, then limits to dynamical age estimates for these galaxies from their profile shapes are comparable with the SED ages obtained from the broadband colors. We also present accurate measurements of the sky-background in the HUDF and its associated 1$\\sigma$ uncertainties. ", "introduction": "In the last decade, ground and space based observations of high redshift galaxies have begun to outline the process of galaxy assembly. The details of that process at high redshifts, however, remain poorly constrained. There is increasing support for the model of galaxy formation, in which the most massive galaxies assemble earlier than their less massive counterparts \\citep[e.g.][]{cowi96,guzm97,koda04,mcca04}. A detailed analysis of the `fossil record' of the current stellar populations in nearby galaxies selected from the \\emph{Sloan Digital Sky Survey} \\citep[SDSS;][]{york00} provides strong evidence for this downsizing picture \\citep{heav04,pant07}. The increasing number of luminous galaxies spectroscopically confirmed to be at $z\\!\\simeq\\!6.5$ \\citep[e.g.][]{hu02, kodi03, kurk04, rhoa04, ster05, tani05}, or \\cle0.9 Gyr after the Big Bang, also supports this general picture. In an alternate hierarchical scenario, arguments have been made that significant number of low luminosity dwarf galaxies were present at these times, and were the main contributor to finish the process of reionization of the intergalactic medium \\citep{yan04a,yan04b}. However, there is presently little information on the dynamical structure of these or other galaxies at $z\\!\\simeq\\!6$. It is not clear whether these objects represent isolated disk systems, or collapsing spheroids, mergers or other dynamically young objects. \\citet{ravi06} used deep, multi-wavelength images obtained with the \\emph{Hubble Space Telescope} (\\emph{HST}) Advanced Camera for Surveys (ACS) as part of the Great Observatories Origins Deep Survey (GOODS) to analyze 2-D surface brightness distributions of the brightest Lyman Break Galaxies (LBGs) at $2.5\\!<\\!z\\!<\\!5$. They distinguish various morphologies based on the S\\'{e}rsic index $n$, which measures the shape of the azimuthally averaged surface brightness profile (where $n$=1 for exponential disks and $n$=4 for a de Vaucouleurs law). \\citet{ravi06} find that 40\\% of the LBGs have light profiles close to exponential, as seen for disk galaxies, and only $\\sim$30\\% have high $n$, as seen in nearby spheroids. They also find a significant fraction ($\\sim$30\\%) of galaxies with light profiles \\emph{shallower} than exponential, which appear to have multiple cores or disturbed morphologies, suggestive of close pairs or on-going galaxy mergers. Distinction between these possible morphologies and, therefore, a better estimate of the formation redshifts of the systems observed at $z\\!\\simeq\\!4\\!-\\!6$ in particular, is important for testing the galaxy assembly picture, and for the refinement of galaxy formation models. One possible technique involves the radial surface brightness profiles of the most massive objects --- those that will likely evolve to become the massive elliptical galaxies, which we see in place at redshifts $z\\!\\lesssim\\!2$ \\citep{driv98,vand03,vand04}. This can be analytically understood in the context of the \\citet{lynd67} relaxation formalism and the numerical galaxy formation simulations of \\citet{vana82}, which describe collisionless collapse and violent relaxation as the formation mechanism for elliptical galaxies. As the time-scale for relaxation is shorter in the inner than in the outer parts of a galaxy, convergence toward a $r^{1/n}$-profile will proceed from the inside to progressively larger radii at later times. Moreover, \\citet{korm77} has shown that tidal perturbations due to neighbors can cause the radial surface brightness profile to deviate from a pure de~Vaucouleurs profile in the outer parts of a galaxy. This implies that the radius where surface brightness profiles start to deviate significantly from an $r^{1/n}$ profile \\emph{might} serve as a ``\\emph{virial clock}'' that traces the time since the onset of the last major merger, accretion events or global starburst in these objects. Image stacking methods have been used extensively on X-ray \\citep{nand02,bran01} and radio \\citep{geor03,whit07} data to study the mean properties (e.g. flux, luminosity) of well-defined samples of sources that are otherwise too faint to be detected individually. \\citet{pasc96} applied such a stacking method to a large number of optically very faint, compact objects at $z\\!=\\!2.39$ to trace their ``average'' structure. This approach was also applied by \\citet{zibe04} to detect the presence of faint stellar halos around disk galaxies selected from the SDSS. An attempt to apply this technique to high redshift galaxies in the Hubble Deep Field \\citep[HDF;][]{will96} was not conclusive (H. Ferguson; private communication) due to the poorer spatial sampling and shallower depth of the HDF compared to the Hubble Ultra Deep Field \\citep[HUDF;][]{beck06}. In this paper, we use the exceptional depth and fine spatial sampling of the HUDF to study the potential of this image stacking technique, and will estimate limits to dynamical ages of faint, young galaxies at $z\\!\\simeq\\!4\\!-\\!6$. The HUDF reaches $\\sim$1.5~mag deeper than the equivalent HDF exposure in the \\ii-band, and has better spatial sampling than the HDF. The HUDF depth also allows us to characterize the sky background very accurately, which is critical for successfully using a stacking method to measure the mean surface-brightness profiles for these faint young galaxies. This paper is organized as follows: In \\secref{observations} we summarize the HUDF observations, and in \\secref{sample} we discuss the selection of our $z\\!\\simeq\\!4,5$ and $6$ samples. In \\secref{analysis} we describe our data analysis, which includes accurately measuring the 1$\\sigma$ sky-subtraction error, the image stacking method to generate mean surface-brightness profiles, and our test of its reliability. In \\secref{results} we present and discuss our results in terms of the average surface-brightness profiles of $z\\!\\simeq\\!4\\!-\\!6$ galaxies, and in \\secref{conclusion} we conclude with a summary of our results. Throughout this paper we refer to the \\emph{HST}/ACS F435W, F606W, F775W, and F850LP filters as the \\bb-, \\vv-, \\ii-, and \\zz-bands, respectively. We assume a \\emph{Wilkinson Microwave Anisotropy Probe} (WMAP) cosmology of $\\Omega_m$=0.24, $\\Omega_{\\Lambda}$=0.76 and \\Ho=73~km~s$^{-1}$~Mpc$^{-1}$, in accord with the most recent 3-year WMAP results of \\citet{sper07}. This implies a current age for the Universe of 13.65~Gyr. All magnitudes are given in the AB system \\citep{oke83}. ", "conclusions": "\\label{results} \\figref{fig8} shows that the mean surface brightness profiles deviate significantly from an inner $r^{1/n}$ profile at radii $r\\!\\gtrsim$0\\arcspt 27--0\\arcspt 35, depending somewhat on the redshift bin. These deviations appear real, with the break/point of departure located \\cge 1.5--2~mag above the 1$\\sigma$ sky-subtraction error and above the PSF-wings. In the following, we discuss several possible explanations for the observed shapes of our composite surface brightness profiles. \\subsection{Galaxies with Different Morphologies} Our test on nearby galaxies (\\figref{fig9}) shows that, if we stack many galaxies with different morphologies (early-type, late-type or spiral galaxies), it is possible to get a slope-change (`break') in the average surface brightness profile. \\citet{ravi06} find that 40\\% of the brighter LBGs at $2.5\\!<\\!z\\!<\\!5$ have light profiles close to exponential, as seen for disk galaxies, and only $\\sim$30\\% have high $n$, as seen in nearby spheroids. They also find a significant fraction ($\\sim$30\\%) of galaxies with light profiles \\emph{shallower} than exponential, which appear to have multiple cores or disturbed morphologies, suggestive of close pairs or on-going galaxy mergers. Therefore, if galaxies at $z\\!\\simeq\\!4\\!-\\!6$ have a variety of morphological types, then the shape of the average surface brightness profile that we see may be due to the stacking of different types of galaxies. Therefore, we find that the exponential and the flatter profiles found by \\citet{ravi06} for galaxies at $2.5\\!<\\!z\\!<\\!5$ also apply to higher redshifts ($z\\!\\ge\\!5$). Also, we believe that it is more likely that the high redshift, faint galaxy population consists primarily of small galaxies with late-type morphologies and with sub-L$^{*}$ luminosities, as seen at $z\\!\\simeq\\!2\\!-\\!3$ \\citep{driv95,driv98}. So if the $z\\!\\simeq\\!4\\!-\\!6$ population consists of such a late-type galaxy population, then the slope-change in the light profiles is likely not the result of co-adding images of objects with disparate morphological types. \\subsection{Central Star Formation/Starburst} \\emph{HST} optical images of galaxies at $z\\!\\simeq\\!4\\!-\\!6$ sample their rest-frame UV ($\\sim$1200 \\Ang), where the contribution from the actively star-forming regions (very young, massive stars) dominates the UV-light. \\citet{hath07} have shown that galaxies at $z\\!\\simeq\\!5\\!-\\!6$ are high redshift starbursts and these galaxies have similar starburst intensity limit as local starbursting galaxies. Therefore, it is possible that galaxies at $z\\!\\simeq\\!4\\!-\\!6$ have centrally concentrated star formation or starburst. This possibility is based on three key assumptions: (1) most of the galaxies at $z\\!\\simeq\\!4,5,6$ are intrinsically later-type galaxies \\citep{driv98,stei99}; (2) the Spectral Energy Distribution (SED) of these galaxies at $z\\!\\simeq\\!4,5,6$ are dominated by early A- to late O-type stars, respectively; and (3) there are no old stars with ages at $z\\!\\simeq\\!4\\!-\\!6$ greater than 2-1 Gyr in WMAP cosmology, respectively. \\citet{hunt06} studied azimuthally averaged surface photometry profiles for large sample of nearby irregular galaxies. They find some galaxies have double exponentials that are steeper (and bluer) in the inner parts compared to outer parts of the galaxy. \\citet{hunt06} discuss that this type of behavior is expected in galaxies where the centrally concentrated star formation or starburst steepens the surface brightness profiles in the center. If that is the case, then one might expect a better correlation between the break in the surface brightness profiles and changes in color profiles. Unfortunately, for our sample of galaxies at $z\\!\\simeq\\!4\\!-\\!6$, we don't have high- resolution restframe \\uu\\bb\\vv color information. The objects are generally too faint for \\emph{Spitzer Space Telescope}, and hence we cannot confirm or reject this possibility for the shape of our composite surface brightness profiles. \\subsection{\\boldmath {Limits to Dynamical Ages for $z\\!\\simeq\\!4,5,6$ Objects}}\\label{ages} The average compact $z\\!\\simeq\\!4\\!-\\!6$ galaxy is clearly extended with respect to the ACS PSFs (\\figref{fig8}), and is best fit by an exponential profile ($n\\!<\\!2$) out to a radius of about $r\\!\\simeq$0\\arcspt 35, 0\\arcspt 31, and 0\\arcspt 27 at $z\\!\\simeq\\!4,5$ and $6$, respectively. The apparent progression with redshift is noteworthy. The radius at which the profile starts to deviate from $r^{1/n}$ (in this case at radius $r$\\cge 0\\arcspt 35--0\\arcspt 27) may be an important constraint to the dynamical time scale of the system, as discussed in \\secref{introduction}. If this argument is valid, then we can estimate limits to the dynamical ages of $z\\!\\simeq\\!4,5,6$ galaxies as follows. In WMAP cosmology, a radius of $r$\\cge 0\\arcspt 35 at $z\\!\\simeq\\!4$ corresponds to $r$\\cge 2.5 kpc. The dynamical time scale \\citep[e.g.,][]{binn87}, $\\tau_{dyn}$, goes as $\\tau_{dyn}$ = $C r^{3/2} \\!/\\!\\sqrt{G\\,M}$, where the constant $C=\\pi/2$. For a typical dwarf galaxy mass range of $\\sim10^9\\!-\\!10^8$\\Msun inside $r$=2.5 kpc, we infer that the limits to the dynamical age would be $\\tau_{dyn}\\simeq$ 90--290 Myr, which is the lifespan expected for a late-type B-star. This means that the last major merger that affected this surface brightness profile and that triggered its associated starburst may have occurred $\\sim$0.20 Gyr before $z\\!\\simeq\\!4$, ---assuming that the star-formation wasn't spontaneous, but associated with some accretion or a merging event. Table 2 shows the break-radius and inferred limits to dynamical ages for the $z\\!\\simeq\\!4\\!-\\!6$ objects. At $z\\!\\simeq\\!5$, we find that the limits to dynamical age at the break radius would be $\\tau_{dyn}\\simeq$ 70--210 Myr, which is the lifespan expected for a mid B-star, while at $z\\!\\simeq\\!6$, $\\tau_{dyn}\\simeq$ 50--150 Myr, which is the lifespan expected for a late O--early B-star. This means that the last major merger that affected these surface brightness profiles at $z\\!\\simeq\\!5$ and $6$ and that triggered its associated starburst may have occurred $\\sim$0.14 and $\\sim$0.10 Gyr before $z\\!\\simeq\\!5$ and $6$, respectively. \\begin{deluxetable}{cccc} \\tablewidth{0pt} \\tablecaption{Dynamical Ages for $z\\!\\simeq\\!4-6$ objects in the HUDF\\label{table2}} \\tablenum{2} \\tablehead{\\colhead{Redshift} & \\colhead{``Break'' Radius$^a$} & \\colhead{``Break'' Radius$^b$} & \\colhead{Dynamical Age$^c$} \\\\ \\colhead{$z$} & \\colhead{(arcsec)} & \\colhead{(kpc)} & \\colhead{($\\tau_{dyn}$)} } \\startdata 4 & 0.35 & 2.5 & 0.09--0.29 Gyr \\\\ 5 & 0.31 & 2.0 & 0.07--0.21 Gyr \\\\ 6 & 0.27 & 1.6 & 0.05--0.15 Gyr \\\\ \\enddata \\tablenotetext{a}{From composite surface brightness profiles (\\figref{fig6} and \\figref{fig8}).} \\tablenotetext{b}{Radius in kpc corresponding to radius in arcsec at given redshift.} \\tablenotetext{c}{If ``break radius'' interpreted as indicator of dynamical age.} \\end{deluxetable} The dynamical time is a lower limit to the actual time available, since it assumes matter starts from rest. Any angular momentum at start will increase the available time. The best-fit SED age from the GOODS \\emph{HST} and \\emph{Spitzer} photometry on some of the brighter of these objects --- using \\citet{bruz03} templates --- is in the range of about $\\sim$150--650 Myr \\citep{yan05,eyle05,eyle07}, the lower end of which is consistent with our limits to their dynamical age estimates, while the somewhat larger SED ages could also be affected by the onset of the AGB in the stellar population increasing the observed \\emph{Spitzer} fluxes and hence possibly overestimating ages \\citep{mara05}. Our age estimates for $z\\!\\simeq\\!4\\!-\\!6$ are consistent with the trend of SED ages suggested for $z\\!\\simeq\\!7$ \\citep{labb06}. It is noteworthy that, given the uncertainties, the two independent age estimates are consistent. If our limits to dynamical age estimates for the image \\emph{stacks} are thus valid, they are consistent with the SED ages, and point to a consistent young age for these objects. Furthermore, the presence of young, massive late O--early B-stars at $z\\!\\simeq\\!6$ has implications for the reionization of the universe. From observations of the appearance of complete Gunn-Peterson troughs in the spectra of $z\\!\\ga\\!5.8$ quasars \\citep{fan06}, we know that the epoch of reionization had ended by $z\\!\\simeq\\!6$. From the steep ($\\alpha$=--1.8) faint-end slope of the luminosity function of $z\\!\\simeq\\!6$ galaxies, \\citet{yan04a,yan04b} concluded that dwarf galaxies, and not quasars, likely finished reionization by $z\\!\\simeq\\!6$. Should the present interpretation of their light profiles be correct, then it would appear to add support to this picture, in the sense that such objects are dominated by B-stars and did not start their most recent major starburst long before $z\\!\\simeq\\!6$." }, "0710/0710.4909_arXiv.txt": { "abstract": "The origin of galactic cosmic rays is one of the most interesting unsolved problems in astroparticle physics. Experimentally, the problem is attacked by a multi-disciplinary effort, namely by direct measurements of cosmic rays above the atmosphere, by air shower observations, and by the detection of TeV $\\gamma$ rays. Recent experimental results are presented and their implications on the contemporary understanding of the origin of galactic cosmic rays are discussed. ", "introduction": "The Earth is permanently exposed to a vast flux of highly energetic particles, fully ionized atomic nuclei from outer space. The extraterrestrial origin of these particles has been demonstrated by V. Hess in 1912 \\cite{hess} and he named the particles \"H\\\"ohenstrahlung\" or \"Ultrastrahlung\". In 1925 R. Millikan coined the term \"Cosmic Rays\". They have a threefold origin. Particles with energies below 100~MeV originate from the Sun. Cosmic rays in narrower sense are particles with energies from the 100~MeV domain up to energies beyond $10^{20}$~eV. Up to several 10~GeV the flux of the particles observed is modulated by the 11-year cycle of the heliospheric magnetic fields. Particles with energies below $10^{17}$ to $10^{18}$~eV are usually considered to be of galactic origin. A proton with an energy of $10^{18}$~eV has a Larmor radius $r_L=360$~pc in the galactic magnetic field ($B\\approx3$~$\\mu$G). This radius is comparable to the thickness of the galactic disc and illustrates that particles at the highest energies can not be magnetically bound to the Galaxy. Hence, they are considered of extragalactic origin. The energy density can be inferred from the measured differential energy spectrum $dN/dE$ \\cite{halzenrhoe} \\begin{equation} \\rho_E=\\frac{4\\pi}{c}\\int \\frac{E}{\\beta} \\frac{dN}{dE} dE , \\end{equation} where $\\beta c$ is the velocity of particles with energy $E$. For galactic cosmic rays the major contribution to the total energy density originates from particles with energies around 1~GeV. Such particles are strongly influenced by the heliospheric magnetic fields. Outside the heliosphere, in the local interstellar environment an energy density $\\rho_E^{LIS}=1.1$~eV/cm$^3$ is obtained (\\rref{gaisserbuch} p.\\,12). The parameterization of the measured galactic cosmic-ray flux according to the \\modell \\cite{pg} results in a density $\\rho_E^{gal}=0.43$~eV/cm$^3$. This is the measurable energy density at Earth (for an average modulation parameter $M=750$~MeV, see (1) in \\rref{pg}). This implies that less than half of the energy flux can be registered directly at Earth. In this article we will give an overview on recent experimental results and their implications on the contemporary understanding of the origin of galactic cosmic rays. As space is limited here, the reader may also consider further recent reviews by the author \\cite{pg,origin,cospar06,vulcano,ecrsreview}. Progress in the understanding of the origin of galactic cosmic rays emerged mainly from observations in three complementary disciplines. The direct measurement of cosmic-ray particles above the atmosphere in outer space and on stratospheric balloons (\\sref{direct}). At energies exceeding $10^{15}$~eV the steeply falling cosmic-ray energy spectrum requires experiments with large detection areas exposed for long times, at present, only realized in ground based installations (\\sref{indirect}). With such detectors extensive air showers are detected, which originate in interactions of high-energy particles in the atmosphere (\\sref{eas}). And, finally, the observation of TeV $\\gamma$-rays (\\sref{gamma}). ", "conclusions": "\\label{outlook} In the last decade our understanding of the origin of galactic cosmic rays has been significantly improved by multidisciplinary efforts, combining key observations of the direct and indirect measurements of cosmic rays as well as the detection of $\\gamma$-rays. The observations by the HESS $\\gamma$-ray telescope give clear hints that hadronic particles are accelerated in SNR. The data are compatible with a model of first order Fermi acceleration at strong shock fronts. The particles propagate in a diffusive process through the Galaxy. Parameters of the propagation models have been constraint by direct measurements above the atmosphere. The KASCADE experiment has shown that the energy spectra of the light elements exhibit a cut-off structure, while the spectra of heavier elemental groups follow power laws to higher energies. The observed spectra seem to be compatible with the assumption of power laws and a cut-off energy proportional to the nuclear charge. This implies that the knee in the all-particle energy spectrum is caused by a cut-off of the light elements. The shape of the all-particle spectrum at higher energies is then determined by the subsequent cut-offs of all elemental species in cosmic rays. Most likely, the astrophysical origin of the knee is a combination of the maximum energy reached in the acceleration process and leakage from the Galaxy during propagation \\cite{prop}. More exotic ideas about the cause of the knee are most likely excluded \\cite{cospar06,ecrsreview}. In conclusion this gives a qualitative 'standard picture' of the origin of galactic cosmic rays. However, several details remain unclear and a precise quantitative description of all aspects of the acceleration and propagation mechanisms is still missing. Among the open questions are: \\\\ It is not clear how to precisely match the spectral indices observed at Earth to the spectra at the sources, being compatible with the TeV $\\gamma$-ray observations and the modification of the spectral slope during propagation.\\\\ A precise astrophysical interpretation of air shower data is at present limited by the understanding of hadronic interactions in the atmosphere, thus the exact shape of the energy spectra at their corresponding knees is unknown.\\\\ Contemporary assumptions on the parameters of cosmic-ray propagation models yield anisotropies in the arrival directions not observed by air shower experiments \\cite{ptuskinaniso,prop}. In the near future experiments like TRACER aim to reveal details of the cosmic-ray propagation for energies approaching the knee \\cite{muellermerida}. Experiments at the LHC probing the extreme forward direction of phase space will improve the description of high-energy hadronic interactions \\cite{engelpylos}. Also the exploration of the end of the galactic component and the transition to extragalactic particles in the energy range from $10^{17}-10^{18}$~eV will be of importance. Key experiments in this energy region are KASCADE-Grande (a 0.5~km$^2$ extension of the KASCADE experiment) \\cite{grande}, Ice Cube/ Ice Top (a 1~km$^2$ air shower experiment and neutrino telescope at the South Pole) \\cite{icecube,icetop}, and HEAT/AMIGA (a 25~km$^2$ extension of the Auger experiment to lower energies) \\cite{klagesmerida}. Around $10^{18}$~eV two features appear in the all-particle energy spectrum. The second knee at $E_{2nd}\\approx400$~PeV$\\approx92\\times E_k$, where the spectrum exhibits a steepening to $\\gamma\\approx-3.3$, and the ankle at about 4~EeV, above this energy the spectrum seems to flatten again to $\\gamma\\approx-2.7$. The region around 4~EeV is sometimes also called \"the dip\" in the spectrum.\\\\ A possible cause for the second knee is the end of the galactic component, when all elements successively have reached their cut-off energies, the latter being proportional to their nuclear charge \\cite{pg,prop}. If one assumes that ultra-heavy elements (heavier than iron) play an important role to understand the second knee, the factor of 92 between the energies of the knee and the second knee can be easily understood as the nuclear charge of the heaviest elements in the periodic table.\\\\ The dip is proposed to be caused by electron-positron pair production of cosmic rays on photons of the cosmic microwave background \\cite{berezinskydip}. The investigation of the transition region and a precise measurement of the galactic all-particle spectrum is also important for an estimate of the energy content of extragalactic cosmic rays. For example, the extragalactic component needed according to the \\modell to sustain the observed all-particle flux at highest energies has an energy density of $\\rho_E^{exg}=3.7\\cdot10^{-7}$~eV/cm$^3$. A promising rediscovered technique for the exploration of cosmic rays from the transition region to highest energies is the detection of radio signals from air showers \\cite{allanrev}. Most likely the emission mechanism is coherent synchrotron radiation of electrons with energies around the critical energy (85~MeV) deflected in the magnetic field of the Earth (geosynchrotron radiation) \\cite{huegefalcke}. The LOPES experiment, registering showers in coincidence with the KASCADE-Grande experiment has demonstrated the feasibility of this approach \\cite{radionature}. Radio detection of air showers is also pursued in the CODALEMA experiment \\cite{codalema} and within the LOFAR radio telescope \\cite{lofar}. Also first radio pulses from air showers have been recorded with antennae set up at the southern site of the Pierre Auger Observatory \\cite{vdbergmerida}." }, "0710/0710.3835_arXiv.txt": { "abstract": "We derive a simple consistency relation from the running of the tensor-to-scalar ratio. This new relation is first order in the slow-roll approximation. While for single field models we can obtain what can be found by using other observables, multi-field cases in general give non-trivial contributions dependent on the geometry of the field space and the inflationary dynamics, which can be probed observationally from this relation. The running of the tensor-to-scalar ratio may be detected by direct laser interferometer experiments. ", "introduction": " ", "conclusions": "" }, "0710/0710.4412_arXiv.txt": { "abstract": "We investigate numerically the long-time behavior of balanced Alfv\\'en wave turbulence forced at intermediate scales. Whereas the usual constant-flux solution is found at the smallest scales, two new scalings are obtained at the forcing scales and at the largest scales of the system. In the latter case we show, in particular, that the spectrum evolves first to a state determined by Loitsyansky invariant and later a state close to the thermodynamic equipartition solution predicted by wave turbulence. The astrophysical implications for galactic magnetic field generation are discussed. ", "introduction": "Turbulence flows are ubiquitous in astrophysical environments from the solar wind (Goldstein et al., 1999; Galtier, 2006), to interstellar (Elergreen et al., 2004; Scalo et al., 2004), galactic and even intergalactic media (Govoni et al., 2006). At the larger scales, signatures of astrophysical turbulence are found, in particular, in the magnetic field measurements whose origin remains one of the major challenging problems (Pouquet, 1993; Widrow, 2002; Brandenburg, 2005). In this paper, we emphasize a new mechanism for generating large-scale magnetic field. It is based on the resonant interactions of shear-Alfv\\'en waves in a turbulent medium permeated by a strong external magnetic field and influenced by an external forcing at intermediate scales. In this situation, both large-scale kinetic and magnetic energies may be produced by mainly non local interactions. This scenario, although very simple, may be relevant for describing re-generation (i.e. maintenance) of large-scale galactic fields, e.g. in our galaxy where energy is injected at intermediate scales by stellar winds and supernovae explosions on scales $10-100$pc (Ferriere et al., 2004). ", "conclusions": "We have seen that the large-scale magnetic field is produced on a time-scale which is about $10^4$ larger than the time-scale needed to establish the small-scale energy cascade solution. We believe that the final distribution in the large-scale part should correspond to the thermodynamic energy equipartition $k_\\perp^{1}$ even though it would take an extremely long computing time for formation of this spectrum which we were not able to achieve. Instead, we observed formation of a shallower long-term scaling at the large scales, $k_\\perp^{0.6}$, which we believe to be transient. Because of the energy cascade to the smallest dissipative (resistive and viscous) scales, the small-scale part of the spectrum is important for understanding the total energy dissipation rate. On the other hand, it is this large-scale part of the spectrum that contains most of the wave (magnetic and kinetic) energy itself. One can view it as a sort of powerful energy storage which is charged extremely slowly and in a very inefficient way (because most of the charging energy is wasted via the energy cascade). For example, if the forcing is concentrated in a thin spherical shell with wavenumber lengths between $k$ and $k + \\Delta k$ then the final energy at large scales (assuming the energy equipartition) will be $k/\\Delta k$ times greater than the energy at the forcing scales. The mechanism of generation of large-scale magnetic fields via nonlinear transfer from an energy source at smaller scales may be relevant, in a very qualitative sense, to maintenance of large-scale galactic fields by small-scale sources provided by supernovae events. To pursue this line further, however, one would have to consider a more realistic thin disk geometry, instead of a simple homogeneous external field considered in the present paper, which could be an interesting future project." }, "0710/0710.4138_arXiv.txt": { "abstract": "We present a detailed analysis of week-long simultaneous observations of the blazar \\mrk at 2--60~keV \\xrays (\\rxtenosp) and TeV \\grays (Whipple and HEGRA) in 2001. Accompanying optical monitoring was performed with the Mt. Hopkins 48\" telescope. The unprecedented quality of this dataset enables us to establish the existence of the correlation between the TeV and \\xray luminosities, and also to start unveiling some of its characteristics, in particular its energy dependence, and time variability. The source shows strong variations in both \\xray and \\gray bands, which are highly correlated. No evidence of a \\xraynosp/\\gray interband lag $\\tau$ is found on the full week dataset, with $\\tau\\lesssim3$\\,ks. A detailed analysis of the March 19 flare, however, reveals that data are \\textit{not} consistent with the peak of the outburst in the 2--4\\,keV \\xray and TeV band being simultaneous. We estimate a $2.1\\pm0.7$\\,ks TeV lag. The amplitudes of the \\xray and \\gray variations are also highly correlated, and the TeV luminosity increases more than linearly with respect to the \\xray one. The high degree of correlation lends further support to the standard model in which a unique electrons population produces the \\xrays by synchrotron radiation and the \\gray component by inverse Compton scattering. However, the finding that for the individual best observed flares the \\gray flux scales approximately quadratically with respect to the \\xray flux, poses a serious challenge to emission models for TeV blazars, as it requires rather special conditions and/or fine tuning of the temporal evolution of the physical parameters of the emission region. We briefly discuss the astrophysical consequences of these new findings in the context of the competing models for the jet emission in blazars. ", "introduction": "\\label{sec:intro} \\object[Mrk 421]{\\mrk} is the brightest BL Lac object in the \\xray and UV sky and the first extragalactic source detected at TeV energies \\citep{punch92_tev_mkn421}. Like most blazars, its spectral energy distribution shows two smooth broad band components (\\eg \\citealp*{sambruna96_sed,umu97_review}; \\citealp{fossati98_sed}). The first one extends from radio to \\xrays with a peak in the soft to medium \\xray range; the second one extends up to the GeV to TeV energies, with a peak presumed to be around 100~GeV. The emission up to \\xrays is thought to be due to synchrotron radiation from high-energy electrons, while the origin of the luminous \\gray radiation is more uncertain. Possibilities include inverse Compton scattering of synchrotron (synchro self-Compton, SSC) or ambient photons (external Compton, EC) off a single electron population thus accounting for the spectral ``similarity'' of the two components (\\eg \\citealp{macomb95_mkn421,mastichiadis_kirk97,tavecchio_tev_98,% mgc92_3c279,sbr94_ec,dermer_etal92}). Alternative ``hadronic'' models produce \\gray from protons, either directly (proton synchrotron) or indirectly (\\eg synchrotron from a second electron population produced by a cascade induced by the interaction of high-energy protons with ambient photons) \\citep{mucke_etal_2003_SPB_model,bottcher_reimer_2004}. The synchrotron proton scenario may be more favorable for objects like \\mrk \\citep{mucke_etal_2003_SPB_model}, because of the lower density of the diffuse photon fields necessary for processes like pion photoproduction to be effective. Moreover, it is generally true that ``hadronic'' models need a higher level of tuning in order to reproduce the observed highly correlated \\xraynosp/\\gray variability. Hence, in this paper we have not addressed this class of models, and instead focused our limited modeling effort on the pure SSC model. All of the above models of the \\gray emission from blazars have all had some degree of success in reproducing both single-epoch spectral energy distributions and their relative epoch-to-epoch changes (\\citealp{vonmontigny95_egret}, \\citealp{gg98_sedt}). These favor the SSC model in \\mrk because it is a BL Lac object for which the ratio of thermal (accretion disk and broad line region) and synchrotron photons is $\\sim0.1$, indicating that the EC mechanism is not important. Detailed modeling of ``blue'' BL~Lacs finds that one-component SSC model can generally account for the time-averaged spectral energy distributions \\citep{gg98_sedt}. Some data sets seem to require modifications of the simple model, introducing either multiple SSC components or additional external seed photons (\\eg \\citealp{blazejowski_etal_2005}). We can, however, further decrease the degeneracy among proposed physical models by taking advantage of blazars' rapid, large-scale time variability with simultaneous \\xraynosp/TeV monitoring (\\eg \\citealp{tavecchio_tev_98,maraschi99_sax_mkn421,henric01_mrk421_x_tev}). Different models produce emission at a given frequency with particles of different energies, cooling times, and cross sections for different processes (\\eg \\citealp{blumenthal_gould70,coppi_blandford90}), and thus are in principle distinguishable \\citep*{henric02_timedependent}. For example, the SSC model predicts nearly simultaneous variations in both the synchrotron and Compton components (see however \\S\\ref{sec:conclusions}), while other models predict more complicated timing (\\eg \\citealp{umu97_review}). With the possible exception of the \\xray and \\gray data taken on Mrk\\,501, the multiwavelength observations on which we base our inferences, have often undersampled the intrinsic variability timescales \\citep{buckley96_mrk421,petry00_mrk501,2001ApJ...563..569T,maraschi99_sax_mkn421} and lack a sufficiently long baseline to make a quantitative assertion about the statistical significance of a correlation. Recently, there has even been evidence of an ``orphan'' TeV flare for the Blazar 1ES\\,1959+650 \\citep{krawczynski_etal_2004_pks1959}; a transient $\\gamma$-ray event that was not accompanied by an obvious \\xray flare in simultaneous data. For what concerns \\mrknosp, there have been regular multiwavelength campaigns in the last several years, planned with an observing strategy focusing on month-long timescales \\citep{blazejowski_etal_2005,rebillot_etal_2006}, and in turn a relatively sparse time sampling (typically one \\rxte snapshot per night, plus binned \\rxtenosp/ASM light curves). These campaigns showed that \\xray and \\gray brightnesses vary ``in step'' and there is certainly a ``loose'' correlation, and also raised some questions about the need of considering additional components to account for the spectra \\citep{blazejowski_etal_2005}, or very high Doppler factors \\citep{rebillot_etal_2006}. The March 2001 campaign remains the experiment with the highest density coverage at both \\xray and TeV energies, and thus the best dataset to address questions concerning the characteristics of the variability of the two spectral energy distribution (SED) components. Moreover, the brightness state achieved during the week of the observations was unprecedented and it remains unparalleled. Preliminary results were presented in \\citet{jordan01_icrc,fossati03_veritas03}. In this paper we present the summary of the multiwavelength observations, with a particular focus on the correlated \\xraynosp/TeV variability, also including the TeV data taken by the HEGRA (High Energy Gamma Ray Astronomy) telescope. An account of the HEGRA March 2001 observations was published by the HEGRA collaboration \\citep{aharonian02_mrk421_spectral_var}. \\citet{giebels_etal_2007_mkn421} report on additional TeV observations with the CAT observatory (Cerenkok Array at Themis), simultaneous with the \\rxte data, not included in this paper. The \\rxte observations, timing and spectral properties are fully presented by Fossati \\etal (2008, in preparation; hereafter F08). The paper is organized as follows: the relevant information about the \\xray and TeV observations, and data reduction is given in \\S\\ref{sec:data}. The observational findings are presented in \\S\\ref{sec:analysis}. and discussed in \\S\\ref{sec:conclusions} in the context of the synchrotron self--Compton model. Section~\\ref{sec:conclusions} summarizes the conclusions. ", "conclusions": "\\label{sec:conclusions} The correlation between the variations in the \\xray and TeV bands is confirmed with unprecedented detail, supporting the idea that the same electron distribution, in the same physical region, is responsible for the emission in both energy bands. However the details of these findings pose a serious challenge to the emission models. Here we would like to sketch a few selected outstanding issues raised by the correlated variability. We refer to a forthcoming paper for an in-depth analysis comprising more extensive modeling. \\begin{itemize} \\item[$\\bullet$] If modeled within the realm of standard values for magnetic field, Doppler factor, source size, the IC scattering responsible for the observed TeV emission occurs in the Klein-Nishina regime. This means that with \\xray and \\gray observations, although we seem to be observing regions of the spectrum that are very similar to each other for what concerns their position with respect to the SED peaks, we are not tracking the evolution of the same electrons (and photons). The extent of the phase and amplitude correlation of the \\xray and \\gray variations is however remarkable, and this sets broad constraints on the characteristics of the processes responsible/governing the variability, \\eg acceleration/injection of particles, dominant cooling cause. \\item[$\\bullet$] In this context, the observation in K-N regime of a quadratic relationship between synchrotron and IC variations (which would be naturally produced in the Thomson regime because of the effectiveness of self-Compton) constrains the electron spectrum variations to occur over an energy band broad enough to affect also the IC seed photons. Moreover, this variation must be essentially ``achromatic'' (\\ie just a change in normalization), otherwise the extra energy-dependent factor would produce an observable effect. \\item[$\\bullet$] The observation that the flux--flux path of the better observed flares decay follows closely the bursting path introduces a further complication. If the flare decay is governed by the cooling of the emitting electrons we do not expect the quadratic relationship to hold during the decaying phase. In fact, given the energy-dependent nature of synchrotron (and IC) cooling, with $\\tau_{cool}\\sim E_{ph}^{-1/2}$, the $50$\\,eV seed photons cool on a longer timescale, \\eg $\\sim$10 times longer than the timescale for photons observed at $\\gtrsim3$\\,keV. The possibility that also the electrons contributing to the bulk of the TeV emission have lower energy than those observed in \\xraynosp, compounds the problem. This means that during the flare decay the \\xray and $\\gamma$-ray brightnesses should follow something like a linear relationship, because the IC (TeV) emission will just reflect the evolution of the electron spectrum, scattering a ``steady'' seed photon field. Plain radiative cooling does not seem to match these observations. A viable mechanisms explaining the flares evolution should allow the concurrent cooling of a broad portion of the electron distribution. On the other hand, brightness variations are accompanied by large spectral changes, and in most cases they are very suggestive of acceleration --or injection-- of the higher energy end of the electron population. \\item[$\\bullet$] Another recurring discrepancy between data and simple one-zone SSC modeling is that of the TeV spectral shape, which is often harder than model predictions. As we illustrated in the previous section the Klein-Nishina effect plays an important role in this respect. A more careful analysis is warranted, but it is worth noting that one alternative option for addressing this problem is that of considering the effect of additional IC components, off photons external to the blob. \\cite{blazejowski_etal_2005} showed that a multi-component model seems to be required to fit the 2003$-$2004 observations, and it might mitigate the discrepancy in the TeV spectrum. \\cite{gg_ft_mc_2005_structure_jets} discuss the effect of including the effect of the radiation emitted by the putative lower Lorentz factor outer layer of the jet, namely as source of additional seed photons for IC. \\item[$\\bullet$] Exotic scenarios could address some of these issues, namely by allowing the IC scattering to occur in the Thomson regime, and hence the self-Compton to be effective, thus reinstating the close relationship between the photons and electrons tracked by \\xray and \\gray observations. One such scenario that we discussed briefly would call for very high values of the Doppler beaming factor, that would reduce the intrinsic energies at play, higher magnetic field and very small size of the emission region. In fact recently there has been some interest for less conventional modeling of TeV blazars (\\eg \\citealt{henric01_mrk421_x_tev,rebillot_etal_2006} for \\mrknosp). However, high Doppler factor scenarios raise a series of new issues, or re-open some that have been settled for the more traditional model. First of all, the fact that the beaming cone of the radiation emitted by such a fast blob is going to be much narrower has to be reconciled with the population statistics of blazars and radio-galaxies, their putative parent population (\\citealt{up95_review}). One way of doing it would be to imagine that the jet comprises a very large number of small high Lorentz factor blobs, fanning out filling a wider cone, with aperture consistent with the unification statistics. This would constitute a quite radical change in the jet structure, from the current one where emission is thought to come from internal shocks. There are also implications concerning the statistical properties of the variability, which would likely be due to the combination of the bursting of different blobs, most likely uncorrelated. \\end{itemize} The richness and depth of the \\xray and \\gray data of the March 2001 campaign presented in this paper raise the bar for models. The aggregate characteristics illustrated here already challenge the simple traditional SSC model, and the SED-snapshot approach. In order to answer the questions raised by these observations it is of paramount importance to exploit fully the time--axis dimension into the modeling and take a dynamical approach. The data time density and brightness (and so statistics) are unparalleled, enabling time resolved spectroscopy on timescales of the order of the physically relevant ones, hence allowing to model the phenomenology self-consistently minimizing the need (freedom) to make assumptions as to how to connect spectra taken at different times. It is likely that this dataset is going constitute the best benchmark for time dependent modeling for some time, despite the great progress made by ground based TeV atmospheric Cherenkov telescopes in the last few years, because of the difficulty of securing long uninterrupted observations with Chandra and XMM-Newton." }, "0710/0710.4248_arXiv.txt": { "abstract": "The {\\it INTEGRAL\\/} satellite, which studies the Universe in the hard X-ray and soft Gamma-ray domain, has been operational for 5 years now. The X-ray telescopes, which use the coded mask technique, provide unprecedented spectral and imaging resolution. This led to a number of discoveries, such as the distribution of diffuse emission in the Galaxy, the discovery of highly absorbed sources and fast X-ray transients in the Galactic Plane, localization of $\\sim50$ Gamma-ray bursts, and the resolution of the cosmic X-ray background around its peak at 30 keV. About 300 previously known X-ray sources have been detected and in addition more than 200 new sources have been discovered. {\\it INTEGRAL\\/} provides spectra starting at 3 keV and ranging up to several hundred keV. This article gives a brief overview about the major discoveries of {\\it INTEGRAL}. ", "introduction": "ESA's {\\it INTEGRAL\\/} space mission \\cite{INTEGRAL} hosts two major hard X-ray instruments, IBIS and SPI, both coded-mask telescopes. IBIS \\cite{IBIS} provides imaging resolution of 12 arcmin, while SPI \\cite{SPI} is optimized for spectroscopy. Both instruments operate at energies from 15 keV up to several MeV. Co-aligned with these main instruments are the two X-ray monitors JEM-X \\cite{JEMX}, which provides spectra and images in the 3--30 keV band, and the optical camera OMC \\cite{OMC}, which provides photometry in the V filter. {\\it INTEGRAL} was launched on October 17, 2002 from Baikonur into a highly eccentric orbit with a perigee of $9,000$ km and an apogee of $150,000$ km, which avoids as much as possible the Earth's radiation belt and allows for un-interrupted observations of up to 3 days. % ", "conclusions": "" }, "0710/0710.2552_arXiv.txt": { "abstract": "We expand our Bayesian Monte Carlo method for analyzing the light curves of gravitationally lensed quasars to simultaneously estimate time delays and quasar structure including their mutual uncertainties. We apply the method to HE1104--1805 and QJ0158--4325, two doubly-imaged quasars with microlensing and intrinsic variability on comparable time scales. For HE1104--1805 the resulting time delay of $\\Delta t_{AB} = t_A - t_B = 162.2_{-5.9}^{+6.3}$~days and accretion disk size estimate of $\\log(r_s / {\\rm cm}) = 15.7_{-0.5}^{+0.4}$ at $0.2\\mu$m in the rest frame are consistent with earlier estimates but suggest that existing methods for estimating time delays in the presence of microlensing underestimate the uncertainties. We are unable to measure a time delay for QJ0158--4325, but the accretion disk size is $\\log(r_s / {\\rm cm}) = 14.9\\pm0.3$ at $0.3\\mu$m in the rest frame. ", "introduction": "Variability in lensed quasar images comes from two very different sources. Changes in the quasar's intrinsic luminosity are observable as correlated variability between images, while microlensing by the stars in the lens galaxy produces uncorrelated variability. Measurements of the time delays between the lensed images from the correlated variability can be used to study cosmology (e.g. \\citealt{Refsdal1964} and recently \\citealt{Saha2006,Oguri2007}) or the distribution of dark matter in the lens galaxy \\citep[e.g.][]{Kochanek2006,Poindexteretal.2007,Vuissoz2007}. The microlensing variability can be used to study the structure of the quasar, the masses of the stars in the lens galaxy, and the stellar mass fraction near the lensed images \\citep{Schechter2002,Wambsganss2006}. It is now possible to use microlensing to measure the correlation of accretion disk size with black hole mass \\citep{Morgan2007}, the wavelength dependence of the size of the accretion disk \\citep{Poindexter2007} or the differing sizes of the thermal and non-thermal X-ray emission regions \\citep{Pooley2007,Dai2007}. The challenge is that most lensed quasars exhibit both intrinsic and microlensing variability. To measure a time delay, one must successfully model and remove the microlensing variability such that only intrinsic variability remains. If the microlensing variability has a sufficiently low amplitude or long timescale, it can be ignored \\citep[e.g. PG1115-080,][]{Schechter1997}, but this is a dangerous assumption for many systems. \\citet{Eigenbrod2005} found that for an $80$~day delay, adding microlensing perturbations with an amplitude of 5\\% (10\\%) to a light curve increased the uncertainty in the time delay by a factor of 2 (6). Existing time delay analyses for lenses with microlensing \\citep[e.g.][]{Paraficz2006,Kochanek2006,Poindexteretal.2007} depend on the intrinsic and microlensing variability having different time scales. These analyses also require that the microlensing variability can be modeled by a simple polynomial function. This approach will clearly fail if the two sources of variability have similar time scales or if the microlensing variability cannot be easily parameterized. \\begin{deluxetable}{lcccccc} \\tabletypesize{\\scriptsize} \\tablecaption{HST Astrometry and Photometry of QJ0158--4325 and HE1104--1805} \\tablehead{Lens & \\colhead{Component} &\\multicolumn{2}{c}{Astrometry} &\\multicolumn{3}{c}{Photometry}\\\\ \\colhead{} &\\colhead{} &\\colhead{$\\Delta\\hbox{RA}$} &\\colhead{$\\Delta\\hbox{Dec}$} &\\colhead{H=F160W} &\\colhead{I=F814W} &\\colhead{V=F555W} } \\startdata QJ0158--4325 & A &$\\equiv 0$ &$\\equiv 0$ &$16.47\\pm0.03$ &$17.81\\pm0.04$ &$18.10\\pm0.13$\\\\ & B &$-1\\farcs156\\pm0\\farcs003$ &$-0\\farcs398\\pm0\\farcs003$ &$17.27\\pm0.03$ &$18.62\\pm0.11$ &$18.91\\pm0.17$\\\\ & G &$-0\\farcs780\\pm0\\farcs016$ &$-0\\farcs234\\pm0\\farcs006$ &$16.67\\pm0.13$ &$18.91\\pm0.06$ &$20.36\\pm0.18$\\\\ \\hline HE1104--1805 & A &$\\equiv 0$ &$\\equiv 0$ &$15.91\\pm0.01$ &$16.40\\pm0.03$ &$16.92\\pm0.06$\\\\ & B &$+2\\farcs901\\pm0\\farcs003$ &$-1\\farcs332\\pm0\\farcs003$ &$17.35\\pm0.03$ &$17.95\\pm0.04$ &$18.70\\pm0.08$\\\\ & G &$+0\\farcs965\\pm0\\farcs003$ &$-0\\farcs500\\pm0\\farcs003$ &$17.52\\pm0.09$ &$20.01\\pm0.10$ &$23.26\\pm0.27$\\\\ \\enddata \\label{tab:hst} \\end{deluxetable} In this paper, we present a new technique for simultaneously estimating the time delay and structure of lensed quasars that exhibit strong microlensing. In essence, we assume a range of time delays and then determine the likelihood of the implied microlensing variability using the Bayesian Monte Carlo method of Kochanek (2004, see also \\citealt{Kochanek2007}). This allows us to estimate the time delays and the quasar structural parameters simultaneously and include the effects of both phenomena on the parameter uncertainties. We apply the method to the two doubly-imaged lenses HE1104--1805 \\citep{Wisotzki1993} and QJ0158--4325 \\citep{Morgan1999}. While HE1104--1805 has a well-measured time delay \\citep{Poindexteretal.2007}, the amplitude of the microlensing ($\\sim 0.05 \\; {\\rm mag \\; yr^{-1}}$ over the past decade) and the fact that it exhibits variability on the 6 month scale of the time delay suggest that it is close to the limit where microlensing polynomial fitting methods \\citep{Burud2001,Kochanek2006} will break down. QJ0158--4325 clearly shows both correlated and uncorrelated variability, but the polynomial methods cannot reliably produce a time delay estimate. We describe the data and our models in \\S\\ref{sec:data}, our new approach in \\S\\ref{sec:analysis} and the application to the two systems in \\S\\ref{sec:results}. In \\S~\\ref{sec:discussion}, we discuss the results and their limitations. We assume a flat $\\Omega_0=0.3$, $\\Lambda_0=0.7$, $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$ cosmology and that the lens redshift of QJ0158--4325 is $z_l=0.5$. Reasonable changes in this assumed redshift have negligible consequences for the results. ", "conclusions": "\\label{sec:discussion} \\citet{Peng2006} used the width of the \\ion{C}{4}~($\\lambda 1549$\\AA) emission line to estimate the black hole mass $M_{BH}=2.37 \\times 10^9 {\\rm M_\\sun}$ in HE1104--1805 and the width of \\ion{Mg}{2}~($\\lambda 2798$\\AA) emission line to estimate the black hole mass $M_{BH}=1.6 \\times 10^8 {\\rm M_\\sun}$ in QJ0158--4325. Using these black hole masses, the quasar accretion disk size - black hole mass relation of \\citet{Morgan2007} predicts source sizes at 2500\\AA~of $\\log(r_s/{\\rm cm}) = 15.9 \\pm 0.2$ for HE1104--1805 and $\\log(r_s/{\\rm cm}) = 15.2 \\pm 0.2$ for QJ0158--4325. If we scale our current disk size measurements to 2500\\AA~using the $R_\\lambda \\propto \\lambda^{4/3}$ scaling of thin disk theory and assume an inclination angle $\\cos i = 1/2$, we find $\\log(r_s/{\\rm cm}) = 15.9^{+0.4}_{-0.5}$ for HE1104--1805 and $\\log(r_s/{\\rm cm}) = 14.8 \\pm 0.3$ for QJ0158--4325, fully consistent with the predictions of the \\citet{Morgan2007} accretion disk size - black hole mass relation. The mixing of intrinsic and microlensing variability in lensed quasar light curves can be a serious problem for estimating time delays \\citep[e.g.][]{Eigenbrod2005} and previous microlensing analyses have been restricted to lenses with known time delays. In HE1104--1805, which must be close to the limits of measuring time delays in the presence of microlensing, we confirm that the approach of fitting polynomial models for the microlensing works reasonably well. However, the dependence of the delay on the assumed model was a warning sign that the formal errors on the delays were likely to be underestimates, as was recognized by \\citet{Poindexteretal.2007}. In our new, non-parametric microlensing analysis of HE1104--1805 we find a modestly longer delay of $162.2_{-5.9}^{+6.3}$~days that quantifies those concerns. Estimates of the quasar accretion disk size are little affected by these small shifts in the time delay. In QJ0158--4325, the microlensing amplitude is larger relative to the intrinsic variability, and traditional methods for determining delays fail. Our new method also fails to measure a delay, but it does allow us to measure the size of the quasar accretion disk despite the uncertainties in the time delay." }, "0710/0710.0282_arXiv.txt": { "abstract": "A three fluid system describing the decay of the curvaton is studied by numerical and analytical means. We place constraints on the allowed interaction strengths between the fluids and initial curvaton density by requiring that the curvaton decays before nucleosynthesis while nucleosynthesis, radiation-matter equality and decoupling occur at correct temperatures. We find that with a continuous, time-independent interaction, a small initial curvaton density is naturally preferred along with a low reheating temperature. Allowing for a time-dependent interaction, this constraint can be relaxed. In both cases, a purely adiabatic final state can be generated, but not without fine-tuning. Unlike in the two fluid system, the time-dependent interactions are found to have a small effect on the curvature perturbation itself due to the different nature of the system. The presence of non-gaussianity in the model is discussed. ", "introduction": "The problem of determining the evolution of large scale perturbations in a background of a multi-component fluid system is central in modern day cosmology \\cite{Kodama:1985bj, Mukhanov:1990me, Wands:2000dp}. In such a system, the interactions between the different fluids are important in determining the evolution of the curvature perturbation \\cite{Malik:2004tf}. Examples of interacting fluid systems demonstrate the importance of such systems,as most notably reheating at the end of inflation \\cite{Albrecht:1982mp,Den:1984tn,Kripfganz:1985mn,Bastero-Gil:2002xr,Dvali:2003em,Kofman:2003nx} and the curvaton scenario \\cite{Enqvist:2001zp,Lyth:2001nq,Moroi:2002rd,Moroi:2001ct,Lyth:2002my}. In contrast to any single fluid system, in a multi-component system the total curvature perturbation, $\\zeta$, generally evolves whenever the non-adiabatic pressure is non-zero, \\ie when interactions between the fluids exist. Evolution of the primordial large scale curvature perturbation can relax the underlying assumptions on the inflationary scenario. Therefore analysis of multi-component fluid systems may affect our view on the physical settings. In addition to the curvature scenario considered in this paper, natural frameworks for such mechanism exist \\eg within a traditional multiple inflationary scenario \\cite{minf} or a string landscape picture \\cite{cliff, multiverse}. Whether a given scenario can effectively modify the primordial spectrum depends on the exact nature of the system. Recent cosmic microwave surveys have also brought attention into the concept of non-gaussianity \\ie how much the spectrum deviates from gaussian distribution. This is especially important in multifield models of inflation including the curvaton scenario \\cite{Malik:2006pm,Sasaki:2006kq}. The mechanism how energy is transferred between the fluids can be described by different methods, \\eg by a constant interaction \\cite{Malik:2002jb} or by utilizing the so-called sudden decay approximation \\cite{Lyth:2001nq, Lyth:2002my}. In a recent paper \\cite{Multamaki:2006} we considered relaxing the assumptions behind these approximations by allowing for time dependent interactions while evolving the full large-scale perturbation equations. Such an approach can better model the micro physics behind a particular physical framework by allowing one to choose the strength and the time at which the interaction is turned on. In contrast, if the interaction between is modeled with a constant interaction term, the fluid begins to decay (or interact) when its decay width is of the order of the Hubble rate, $\\Gamma\\sim H$. Physical scenarios relevant to having time (and space) dependent interactions include \\eg phase transitions, multiple inflation scenarios \\cite{minf} and scenarios where locally different decay rates of the inflaton are generated by spatially varying reheating temperature and couplings \\cite{Matarrese:2003tk, Dvali:2003em, Kofman:2003nx}. In the present paper we consider the curvaton scenario with time dependent interactions between curvaton and other fluids. During inflation the curvaton is a light scalar field that does not contribute to the expansion of the universe; after the inflaton field has decayed into relativistic particles the curvaton begins to oscillate and to decay into radiation and matter. The focus of this article is on this situation: we study how a time-dependent interaction affects the evolution of the curvature perturbation. We study the physically allowed parameter space by utilizing information from known cosmological epochs. Moreover, we calculate the amount isocurvature in terms of curvaton decay widths. Finally, we also discuss how to describe non-gaussianity in the three fluid model and present the non-linearity factor $f_{NL}$ \\cite{Komatsu:2001rj}. This paper is organized as follows. In sections II and III we present the governing equations of the background and perturbations in the Newtonian gauge, including discussion of non-gaussianity as described by the $f_{NL}$ parameter. In section IV we present our results while the discussion and conclusions are presented in the following section V. ", "conclusions": "The curvaton model has gained a lot of attention in the recent years mainly because it can make the inflation potential look more natural \\cite{Lyth:2001nq}. Since the first model where the curvaton decayed only into radiation \\cite{Lyth:2001nq,Moroi:2001ct}, a number of other possibilities have been explored including a curvaton web model \\cite{Linde:2005yw}, the possibility of multiple curvaton fields \\cite{Assadullahi:2007uw} and different particle models such as axions \\cite{Chun:2004gx,Dimopoulos:2005bp} just to name a few. In the present paper we have studied a three-fluid model in which the curvaton decays into both radiation and cold dark matter. This has been studied previously also in \\cite{Gupta:2003jc,Lyth:2002my,Gordon:2002gv}. We have assumed that the initial system has no matter content and it is dominated by the radiation which originates from the decay of the inflaton field. Because all of the matter content of the universe comes from the curvaton field we are able to estimate the reheating temperature. This can be additionally used to constraint the parameter space of the model. We have systematically scanned the parameter space and identified the regions where the model is physically acceptable, \\ie when evolution during and after nucleosynthesis is standard while requiring that the reheating temperature is not unreasonably high. We have identified these regions both when the decay rates are fixed and when the curvaton starts to decay later. These allowed regions are alike once the rescaling $\\Gamma/H_0 \\rightarrow \\Gamma/H(N_*)$ is taken into account and the real difference appears in the initial system temperature. We find that if the decay rates are comparable to the Hubble rate, a small initial curvaton density is required. Otherwise one needs to fine-tune the decay rates to be much smaller than $H$ at the time of decay. In the continuous interaction case, requiring $\\Gamma_i\\sim H$ leads to a low reheat temperature, but this can be avoided when the matter interaction is delayed. If the initial curvaton density is large, the final state is naturally adiabatic assuming that the system is otherwise physically acceptable. Note however, that this in turn requires fine-tuning in the decay rates. If $\\Gamma_i\\sim H$, we find that the final state generally contains a large isocurvature component. We have also we studied non-gaussianity in the framework of the three-fluid models. We find that in the region where the first-order perturbation theory can be applied, the three-fluid model gives no limits on the $f_{NL}$ parameter. This is the result of a conserved curvature perturbation $\\zeta_c$, which carries the initial curvaton perturbation $\\zeta_{\\sigma}$ into the matter perturbation $\\zeta_m$ \\cite{Gupta:2003jc}. Our results differ from the previous results \\cite{Lyth:2002my,Gupta:2003jc} mainly because our $f_{NL}$ is evaluated at the time of last scattering and not at nucleosynthesis. This allows the non-gaussianity to be more easily compared to the observational Sachs-Wolfe effect and we do not have to use a radiation transfer function. In order to calculate the observational non-gaussianity, second-order perturbation theory needs to be applied \\cite{Bartolo:2004if}. This is however beyond the scope of this article and will be the focus of a follow-up paper. \\subsection*" }, "0710/0710.2622_arXiv.txt": { "abstract": "Lyman-alpha (\\lya) is one of the dominant tools used to probe the star-forming galaxy population at high-redshift ($z$). However, astrophysical interpretations of data drawn from \\lya\\ alone hinge on the \\lya\\ escape fraction which, due to the complex radiative transport, may vary greatly. Here we map the \\lya\\ emission from the local luminous blue compact galaxy Haro\\,11, a known emitter of \\lya\\ and the only known candidate for low-$z$ Lyman continuum emission (LyC). To aid in the interpretation we perform a detailed UV and optical multi-wavelength analysis and model the stellar population, dust distribution, ionising photon budget, and star-cluster population. We use archival X-ray observations to further constrain properties of the starburst and estimate the neutral hydrogen column density. The \\lya\\ morphology is found to be largely symmetric around a single young star forming knot and is strongly decoupled from other wavelengths. From general surface photometry, only very slight correlation is found between \\lya\\ and \\halpha, \\ebv, and the age of the stellar population. Only around the central \\lya-bright cluster do we find the \\lya/\\halpha\\ ratio at values predicted by recombination theory. The total \\lya\\ escape fraction is found to be just 3\\%. We compute that $\\sim 90$\\% of the \\lya\\ photons that escape do so after undergoing multiple resonance scattering events, masking their point of origin. This leads to a largely symmetric distribution and, by increasing the distance that photons must travel to escape, decreases the escape probability significantly. While dust must ultimately be responsible for the destruction of \\lya, it plays little role in governing the observed morphology, which is regulated more by ISM kinematics and geometry. We find tentative evidence for local \\lya\\ equivalent width in the immediate vicinity of star-clusters being a function of cluster age, consistent with hydrodynamic studies. We estimate the intrinsic production of ionising photons and put further constraints of $\\sim 9$\\% on the escaping fraction of photons at 900\\AA. ", "introduction": "Four decades ago \\cite{pp67} discussed the prospects of identifying `primeval' galaxies (i.e. galaxies forming their first generation of stars), using both the Lyman decrement and Lyman-alpha (\\lya) emission line as observational tracers. Both methods rely upon the galaxies hosting numerous young, massive stars producing a strong radiation field in the far ultraviolet (UV). The absorption of this radiation field bluewards of the Lyman absorption edge results in the Lyman-break phenomenon, while the reprocessing of the absorbed photons in astrophysical nebulae results in the superimposition of strong hydrogen recombination lines on the galaxy spectra. While it's likely that both features are present in the spectra of high-redshift ($z$) starbursts, from a survey perspective they compete. Lyman-break candidates are identified in multi-broadband imaging surveys where large ranges in redshift (and therefore cosmic volume) can be probed, but are biased towards the detection of galaxies with FUV continua towards the brighter end of the luminosity function (LF) -- less numerous in the hierarchical galaxy formation scenario. On the other hand, because emission lines concentrate a large amount of energy in a very small spectral region, sources with much fainter continuum can be uncovered. Unfortunately, isolation of a line requires either a spectroscopic surveys which typically probe narrow fields of view because of the slit dramatically reducing the size in one dimension, or narrowband imaging which probes only a small range in redshift. The advent of large diameter reflectors, efficient optics, and sensitive photometric devices has resulted in both techniques enjoying much success in recent years. Young galaxies with low dust contents can be expected to produce \\lya\\ emission with high equivalent widths (\\wlya), $\\sim 240$\\AA\\ for solar-like metallicities ($Z$) and standard initial mass functions (IMF) \\citep{charlotfall93}. This value is predicted to increase significantly as metallicities approach the population {\\sc iii} domain \\citep{schaerer03}. Since the \\lya\\ rest wavelength lies in the FUV, the line is still observable from the ground in the optical domain at $z\\sim 6$ and beyond making it, in principle, the ideal tool by which to identify star-forming galaxies in the early universe. Despite a rocky start \\citep{pritchet93} \\lya\\ emitters (LAEs) are showing up en masse in high-$z$ surveys and perhaps their observed and predicted number-densities are now converging \\citep{ledelliou06}. \\lya\\ is now being used to put constraints on the final stages of cosmic reionisation \\citep{malhotrarhoads04,dijkstra06} and explore high-redshift large scale structure and galaxy clustering \\citep{hamana04,malhotra05,murayama07}. In addition, numerous sources have been detected through narrowband imaging techniques with very high \\wlya \\citep{malhotrarhoads02,shimasaku06}, perhaps indicative of top heavy IMFs or extreme stars, making such targets ideal to search for signatures of the so-far illusive population {\\sc iii} stars. However, so far spectroscopic observations have failed to detect the strong He{\\sc ii}$\\lambda 1640$\\AA\\ feature expected from pop {\\sc iii} objects (eg. \\citealt{dawson04,nagao07}). Finally, \\lya\\ is also being used to estimate the cosmic star formation rate density at the highest redshifts \\citep{fujita03,yamada05}. Such star formation rates (SFR) are typically estimated assuming case B recombination \\citep{brocklehurst71} and the \\halpha\\ SFR calibration of \\cite{kennicutt98}, although the large spread in SFR(\\lya) {\\em vs.} SFR(FUV) and consistent underestimates from SFR(\\lya) (eg. \\citealt{murayama07}) suggest that such a calibration should be used with caution. This caution must be extended to all cosmological studies in light of the fact that only a fraction of UV-selected high-$z$ targets show a \\lya\\ feature in emission \\citep[e.g.][]{shapley03}. The complexities of using \\lya\\ as a cosmological tool result from the fact that it is a resonant line, and its potential cosmological importance has motivated a number of studies of star-forming galaxies at low-$z$ and theoretical models of resonant line radiative transfer. Early studies with the {\\em International Ultraviolet Explorer (IUE)} initially mirrored the first results at high-$z$: \\lya\\ was typically weak or absent in local starbursts. This systematic weakening of \\lya\\ was first attributed to dust absorption (eg. \\citealt{charlotfall91}) and an anticorrelation between \\wlya\\ and metallicity was found in the {\\em IUE} sample \\citep{charlotfall93}. However, the damped \\lya\\ absorption seen in some local starbursts, and the failure of dust corrections to reconcile \\lya\\ with the fluxes predicted by recombination theory \\citep{giavalisco96}, is indicative of selective attenuation of \\lya. This is clearly exemplified by the fact that the most metal-poor galaxies known at low-$z$ ({\\sc i}Zw\\,18 and SBS\\,03350-52) are shown to be damped \\lya\\ absorbers \\citep{kunth94,thuanizotov97} in spectroscopic observations from the {\\em Goddard High Resolution Spectrograph (GHRS)} and {\\em Space Telescope Imaging Spectrograph (STIS)}. This is to be expected if the starburst is enshrouded by a static layer of neutral hydrogen that is able to resonantly trap \\lya, thereby greatly increasing the path-lengths of the photons in order to escape the host, and exponentially increasing the chance of their destruction by dust grains \\citep{neufeld90}. Further {\\em GHRS} observations \\citep{kunth98} showed that when \\lya\\ is seen in emission, it almost systematically shows a P\\,Cygni profile and systematic velocity offset from low ionisation state metal absorption features in the neutral ISM, suggesting that an outflowing medium is an essential ingredient in the formation of the line profile and \\lya\\ escape physics. Similar results have also been found at high-$z$ \\citep{shapley03,tapken07}. Hydrodynamic models of expanding bubbles \\citep{tt99} predicted \\lya\\ line profiles to follow an evolutionary sequence starting with pure absorption at the earliest times, developing through a pure emission phase into P\\,Cygni profiles as the ISM is driven out by mechanical feedback from the starburst, and fading back into absorption at late times. This allowed \\cite{mashesse03} to reconcile the variety of \\lya\\ profiles observed at low-$z$ in such an evolutionary sequence. Further theoretical studies \\citep[e.g.][]{ahn03,verhamme06} have shown how a wide variety of P\\,Cygni-like and asymmetric profiles can develop, depending on the ISM properties and geometry. The line formation becomes more complex still if the ISM is multiphase \\citep{neufeld91,hansen06} and certain physical and kinematic configurations may lead to increases in the fraction of escaping \\lya\\ photons compared to non-resonant radiation, thereby effectively boosting \\wlya. Recent advances have been made in the understanding of observed line profiles with the implementation of full 3-D codes with arbitrary distributions of gas, ionisation, temperature, dust, and kinematics \\citep{verhamme06}. Since nebular ionisation does not occur in situ with ionising sources, recombination line imaging (e.g. \\halpha) may reveal a morphology different from that found by imaging of the stellar continuum. This phenomenon may be much more significant for \\lya\\ photons which resonantly scatter. While \\lya\\ and \\halpha\\ have the same points of origin, it is likely that internal \\lya\\ scattering events may cause \\lya\\ photons to be emitted far from their production sites, and not be spatially correlated with \\halpha\\ or other non-resonant recombination lines. Targeted spectroscopic observations are therefore liable to miss a fraction of the emission and, while containing neither frequency nor kinematic information, \\lya\\ imaging becomes an invaluable complement to the spectroscopic studies. This was the motivation for our imaging survey of local starbursts using the {\\em Advanced Camera for Surveys (ACS)} (\\citealt{kunth03,hayes05,ostlin07} in preparation). This is a truly unique dataset for a number of reasons. Firstly the angular resolution of the {\\em ACS} allows us to map the \\lya\\ morphology on scales of 5--15~pc; 2--3 orders of magnitude better than typical studies at high-$z$. The addition of \\halpha\\ (absent in almost all high-$z$ studies due to an inconvenient rest wavelength) allows us to quantitatively examine the decoupling of \\lya\\ from non-resonant lines and estimate the global escape fractions. Multiband UV and optical data allow us to map dust reddening, stellar ages and masses, ionising photon production, and other properties of the host, all at the same resolution as \\lya. The first detailed \\lya\\ imaging study of a local starburst, ESO\\,338-IG04 \\citep{hayes05} found emission and absorption varying on very small scales in the central starburst regions, and little or no correlation with the FUV morphology. The starburst is surrounded with a large, diffuse, low surface brightness \\lya\\ halo that contributes $\\sim 70$\\% to the global \\lya\\ luminosity, resulting from the resonant decoupling and diffusion of \\lya. The total escape fraction was found to be just $\\sim 5$\\%, implying any global values (eg. SFR) that would be estimated from \\lya\\ alone would be seriously at fault. Feedback processes from star-formation are capable of driving galaxy-scale `superwinds' that shock heat and accelerate the ambient medium and circumnuclear gas, resulting in large-scale, diffuse X-ray nebulae \\citep{heckman90}. Typically, hot X-ray emitting regions exhibit a tight morphological correlation with with warm gas as probed by optical emission lines \\citep{grimes05}. Indeed, such X-ray nebulae are near-ubiquitous in starbursts \\citep{strickland04} and have been calibrated as tracers of star-formation rates \\citep{ranalli03,colbert04}. \\cite{grimes05} have also shown that observed X-ray spectra are well reproduced by a single thermal plasma over a range of galaxy classifications spanning dwarfs, discs, and ultra-luminous infrared galaxies (ULIRG), although the central regions of some ULIRGs require an additional power-law component. Since the diffuse X-ray component can be so valuable for an understanding of the wind properties, and feedback is an essential ingredient in the escape of \\lya\\ \\citep{kunth98}, X-ray information provides a valuable supporting dataset for a detailed investigation of \\lya. In this article we turn our attention to another target in our sample, the well-known, luminous ($M_B$ = -20.5), low metallicity ($\\log(O/H)+12=7.9$; \\citealt{bergvall02}) blue compact galaxy (BCG). It exhibits a complex morphology consisting of three main star-forming condensations and an unrelaxed kinematic structure \\citep{ostlin99,ostlin01}, suggestive of a dwarf galaxy merger. It is actively star-forming, exhibits a large number of bright young star clusters \\citep{ostlin00}, is a known emitter of \\lya\\ \\citep{kunth98}, and the only known local candidate emitter of Lyman continuum (LyC) \\citep{bergvall06,grimes07}. The FUV continuum luminosity puts Haro\\,11 at the brighter end of the distribution of \\lya\\ emitters at $z\\sim 3.1$ \\citep{gronwall07}. We utilise {\\em HST/ACS} images in the FUV ({\\em Solar Blind Channel; SBC}), to examine \\lya\\ and nearby continuum, and broadband images in the UV and optical, and narrowband \\halpha\\ ({\\em High Resolution Camera; HRC, and Wide Field Camera; WFC}) to examine dust, ages in the stellar population, the star cluster population as a whole, and estimate the LyC production. We use deep ground-based narrowband images in \\halpha\\ and \\hbeta\\ in order to estimate extinction in the gas phase. X-ray observations of Haro\\,11 have been obtained using the {\\em Chandra} satellite \\citep{grimes07} but currently their status is still proprietary. We hence also exploit serendipitous off-axis observations from the {\\em Chandra} and {\\em XMM-Newton} X-ray observatories to study the wind properties and internal photoelectric absorption. The article is arranged in the following manner: in Section~\\ref{sect:data} we describe the observations and data reductions; in Section~\\ref{sect:res} we present the results; in Section~\\ref{sect:andis} we analyse discuss the results; and in Section~\\ref{sect:conc} we present our concluding remarks. We assume a cosmology of $H_0=72$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_\\mathrm{M}=0.3$ and $\\Omega_\\Lambda=0.7$ throughout. The redshift of Haro\\,11 is taken to be 0.020598 from NED, corresponding to a luminosity distance of 87.1~Mpc. \\section[]{Observations, reductions, and data processing}\\label{sect:data} \\subsection{{\\em HST} Ultraviolet and optical} \\subsubsection{Observations} This study makes use of images from all three channels of the {\\em ACS} onboard {\\em HST}: the {\\em SBC} for the FUV; the {\\em HRC} for the 2200\\AA\\ and $U-$band; and the more sensitive {\\em WFC} for images in the optical domain. The details of the observations and post-reduction processing of the {\\em HST} UV and optical dataset are described in a companion paper, \\cite{ostlin07}. Observations were performed in the bandpasses listed in Table~\\ref{tab:exptime}. Briefly, {\\em F122M} and {\\em F140LP} correspond to \\lya\\ on-line and continuum, respectively, while {\\em FR656N} covers \\halpha\\ on-line and {\\em F550M} measures line-free continuum bluewards of \\halpha. The remainder are broadband continuum observations, selected in order to avoid the strongest emission lines. \\begin{table} \\caption{Observations and exposure times} \\centering \\begin{tabular}{@{}lllll@{}} \\hline \\hline Bandpass & &Channel & ExpTime [ s ] & \\# Split \\\\ \\hline {\\em F122M} & \\lya & {\\em SBC} & 9095 & 5 \\\\ {\\em F140LP} &\\lya\\ cont & {\\em SBC} & 2700 & 5 \\\\ {\\em F220W} &$\\sim 2200$\\AA & {\\em HRC} & 1513 & 3 \\\\ {\\em F330W} &$U-$band & {\\em HRC} & 800 & 2 \\\\ {\\em F435W} &$B-$band & {\\em WFC} & 680 & 2 \\\\ {\\em F550M} &medium $\\sim V$ & {\\em WFC} & 471 & 2 \\\\ {\\em FR656N} &\\halpha & {\\em WFC} & 680 & 2 \\\\ {\\em F814W} &$I-$band & {\\em HRC} & 100 & 1 \\\\ \\hline \\end{tabular} \\label{tab:exptime} \\end{table} All images are `drizzled' using the {\\tt MULTIDRIZZLE} task in {\\tt IRAF/STSDAS} onto the same pixel sampling scale (0.025\\arcsec /pixel) and position angle. The inverse variance weight maps were saved from the drizzle process since they provide an estimate of the error in each pixel. Remaining cosmic rays, charge transfer tracks, and blemishes are removed from the CCD observations using the {\\tt CREDIT} task. Remaining band-to-band discrepancies in the astrometric alignment were rectified using the {\\tt GEOMAP} and {\\tt GEOTRAN} tasks. Since our dataset comprises observations between 1200 and 9000\\AA\\ and utilises three different {\\em ACS} channels, we address the issue of variations in the point spread function (PSF) of the images. PSF models of each band were generated using the {\\tt PSF} task in {\\tt DIGIPHOT/DAOPHOT}, with the resulting models being used to convolve all images to the PSF of {\\em F550M} (the broadest emission-line free PSF in our dataset). PSF models were re-computed for the convolved images and compared to that of {\\em F550M}; all frames showed PSF full width at half maximum consistent with {\\em F550M} at below the 5\\% level. \\subsubsection{\\lya\\ continuum subtraction and SED modeling}\\label{sect:contsub} As described in \\cite{hayes05}, and in more detail in Hayes et al. (2007, in preparation), continuum subtraction of \\lya\\ using these filters requires more sophisticated techniques than most emission lines. Because the offline filter is removed from the online filter by $\\Delta \\lambda / \\lambda = 0.22$ and the FUV continuum evolves rapidly as a function of $\\lambda$, age, and \\ebv, it is imperative to understand the behaviour of the continuum between {\\em F140LP} and {\\em F122M}. Special care is also required because of the broad nature of the {\\em F122M} filter which has a rectangular width around 10\\% of the central wavelength, and transmits both the geocoronal \\lya\\ line and Milky Way \\lya\\ absorption profile. In addition, P\\,Cygni profiles may result in a reduction of the net emission due to the blue-side absorption cancelling some or all of the emission. We defined in \\cite{hayes05} the {\\em Continuum Throughput Normalisation} (\\ctn) factor as the factor that, for a given spectrum, scales the continuum flux sampled in {\\em F140LP} to the flux that would be expected from continuum processes alone in {\\em F122M}. The procedure of estimating \\ctn\\ in each pixel utilises each of the images between {\\em F140LP} and {\\em F814W}, the filter throughput profiles, and fitting {\\em Starburst99} spectral evolutionary models \\citep{leitherer99,vazquez05}. In \\cite{hayes05} we demonstrated that we could find non-degenerate best-fitting spectra if SED datapoints sample the UV continuum slope and Balmer/4000\\AA\\ break. That way, for each pixel we fit burst age and \\ebv\\ using standard $\\chi^2$ minimisation. The method has been substantially developed and is described in detail in Hayes et al. (2007, in preparation). We now use continuum-subtracted \\halpha\\ to first map the contribution to the overall SED from nebular gas alone. Thanks to observations at $V$ and $I$ we are also able to constrain the contribution from any stellar population that underlies the current starburst. Essentially we measure and subtract the gas spectrum and fit age and mass in two stellar components, applying the same reddening for all three SEDs. For each pixel, we are able to find the age of the stellar populations and \\ebv, treating the nebular gas and two stellar components independently. With the best-fitting spectral reconstruction (composite {\\em Starburst99} spectra) at each {\\em HST} pixel we then map the \\ctn\\ factor and are able to reliably continuum subtract \\lya. In Hayes et al. (2007, in preparation) we present extensive simulations, designed to test the reliability of this methodology for an array of input spectra. We determine that the method must account for the presence of an underlying stellar population and nebular continuum emission. Both these components may affect the $U-B$ colour and result in bad fits and poor recovery of \\ctn\\ in cases where they dominate the optical luminosity. Our method of reconstructing the nebular continuum and fitting multiple stellar components allows robust recovery of the FUV continuum features and \\ctn. An independent measure of the metallicity is also found to be a requirement although the metallicity of Haro\\,11 is well known. Details of the IMF are found not to have a detrimental impact when multiple stellar component fitting (i.e. a more complex star-formation history) is used. Overall we find that the software employed here is able to always recover input \\lya\\ equivalent widths to within 30\\% for `weak' \\lya\\ emission (\\wlya=10\\AA) and to within 10\\% when the \\lya\\ line is stronger (\\wlya=100\\AA). With regard to the integrated fluxes and visual morphologies, we found in \\cite{hayes05} that varying the parameters of the model spectra had a minimal effect on our results. Morphologies were always indistinguishable by eye and integrated \\lya\\ fluxes self-consistent to within $\\sim 25$\\%, even when pushing the parameters outside the regimes deemed reasonable for the galaxy in question. This procedure also provides numerous other outputs against which we can compare \\lya. The full list of output maps is: continuum subtracted \\lya\\ and \\halpha, \\ctn-factor at \\lya\\ ({\\em F122M/F140LP}), continuum flux densities at 900\\AA\\ (\\flyc), 1500\\AA\\ (\\ffuv), and 2200\\AA\\ (\\fnuv), \\ebv, the age of the two stellar components, the mass of the two stellar components, the stellar equivalent width in absorption of \\halpha\\ and \\hbeta, and $\\chi^2$. These allow us to compare \\lya\\ fluxes and equivalent widths with various local stellar ages, \\ebv, local star-formation rates, and estimate the ionising photon production. Haro\\,11 metallicity is known to be around 20\\% solar \\citep{bergvall02} so for this study we adopt the evolutionary tracks generated with metallicity $Z=0.004$. Since we are concerned with OB-dominated, young starbursts we use the tracks of the Geneva group. In the absence of a quantitative estimate, the IMF was assumed to follow that of Salpeter ($\\alpha=-2.35$) in the range 0.1~M$_\\odot$ to 120~M$_\\odot$. Star-formation history was taken to be that of an instantaneous starburst (single stellar population) with a mass normalisation of $10^6$~M$_\\odot$. \\subsubsection{Binning}\\label{sect:reductionbinning} As described in the introduction, we can expect to see \\lya\\ features (either emission or absorption) around clusters where variations can be expected on small scales, and/or large-scale diffuse emission at low surface brightness. Typically $S/N$ per resolution element is high in the vicinity of strong continuum sources but significantly lower $(< 1)$ in the diffuse regions. Ordinarily, such problems are overcome by smoothing. While fixed kernel smoothing may improve $S/N$ in diffuse regions it smears out spatial details in regions where $S/N$ is good and smoothing would not be necessary. Adaptive smoothing may maintain some spatial resolution but does not necessarily conserve surface brightness. Adaptive binning provides an alternative to the smoothing approach and for this study we make use of the Voronoi tessellation code of \\cite{die06}, a generalisation of the Voronoi binning algorithm of \\cite{cappel03}. Voronoi tessellation overcomes the drawbacks of smoothing by binning together groups of pixels to conglomerate resolution elements (frequently referred to as ``spaxels\"), recomputing the signal and noise in each new bin. Pixels are continuously accreted until a threshold $S/N$ has been met in each spaxel. The advantage of this adaptive binning method is that in high-$S/N$ regions, the spatial sampling remains high because spaxels are typically small, whereas in diffuse regions $S/N$ is improved greatly. Diffuse emission regions, by definition, show variations over much larger spatial scales. Since each individual pixel is used exactly once, surface brightness is always conserved. The {\\em F140LP} FUV science observation is assigned as the ``reference\" image, and used to generate the binning pattern, using the inverse variance map output by the {\\tt STSDAS/MULTIDRIZZLE} task to compute $S/N$. Pixels are accreted into spaxels until $S/N=5$ has been met, with a maximum size of $40^2$pixels $ = 1\\square$\\arcsec\\ in order to speed up the process and prevent spaxels from growing arbitrarily large. \\subsubsection{Super star clusters} Point-like sources are identified in the deepest of the optical bandpasses, {\\em F435W} and {\\em F550M} using the {\\tt DAOFIND} task in {\\tt IRAF}. To be considered a detection, a cluster must be present in both of these bands. Each object was then inspected by eye and any detections that were clearly spurious were removed from the catalogue. For crowded fields PSF-photometry is preferred to standard aperture methods in order to eliminate cross-contamination from neighbouring clusters. However, some clusters may be extended enough to be resolved by our observations, thus precluding the use of PSF-fitting methods and limiting us to aperture photometry. Aperture photometry was performed using the {\\tt PHOT} task in {\\tt IRAF} in all bands using the {\\em F435W+F550M} detected catalogue. An aperture of 0.10\\arcsec\\ was used, and sky was sampled in a circular annulus of radius between 0.1 and 0.15\\arcsec. Aperture corrections were then applied in accordance with those given in \\cite{siri05} for the {\\em HRC} and {\\em WFC} bandpasses. While {\\em SBC/F140LP} aperture corrections have been computed in the past, \\citep[e.g.][]{dieball05}, they are not available in the published literature and our own aperture corrections, computed with the {\\tt TinyTim} software \\citep{tinytim}, are included here in Table~\\ref{tab:apcorr}, Appendix~\\ref{sect:apcorr}. Since we have no a priori knowledge of the continuum slope, we adopt the {\\em F140LP} aperture correction for $\\beta=0$, using 0.1\\arcsec\\ as in the other bandpasses. Since {\\em SBC/F122M} contains the \\lya\\ line, no point-source photometry was performed in this bandpass itself. Instead, we perform aperture photometry on the continuum subtracted \\lya\\ and \\halpha\\ images in the same 0.1\\arcsec\\ apertures at the position of each cluster with no re-centering applied. This gives a direct measure of the \\lya\\ flux and equivalent width in the vicinity of each SSC. In order to estimate the properties of the SSCs, the same SED-fitting software described in Section~\\ref{sect:contsub} was applied to the aperture extracted fluxes. Age, \\ebv, photometric mass, etc. were computed, allowing us to compare these properties with \\lya\\ in the immediate vicinity of each SSC. \\subsection{ESO -- New Technology Telescope} Haro\\,11 was observed during the nights of 18, 19, and 20 September 2004, using the {\\em New Technology Telescope} at ESO La Silla, as part of an observing run to obtain \\halpha, \\hbeta, and [O{\\sc iii}] narrowband images for all southern targets in our {\\em HST} \\lya\\ sample (Atek et al., in preparation). On the night beginning 18 Sept seeing was good, not exceeding 1.2\\arcsec for the duration, although thin cirrus prevents a direct calibration of these data. On the night of 19 Sept, observational conditions were ideal: photometric and with sub-arcsecond seeing throughout. The final night was still photometric but seeing deteriorated to $>2$\\arcsec\\ and images were unusable for science purposes. The good-seeing data from the night of 18 was calibrated through secondary standard stars in the field, using data from the photometric nights of 19 and 20 Sept. Narrowband imaging was performed at \\halpha, \\hbeta, and nearby continuum. Spectrophotometric standard stars LDS749B, Feige 110, and GD50 were selected (on the grounds of their high spectral resolution: $1-2$\\AA, in order to resolve the stellar absorption features around \\halpha\\ and \\hbeta) from the catalog of \\cite{oke90}, and were observed at regular intervals for the duration of each night in all filters. Imaging observations were performed using both the {\\em ESO Multi-Mode Instrument (EMMI)} \\citep{dek86} and {\\em Super Seeing Imager 2 (SuSI2)} \\citep{odor98}, interchangeably. The instruments were used in $2\\times 2$ pixel binning mode to reduce the readout noise, providing a plate-scale of 0.1665\\arcsec~pix$^{-1}$ and 0.130\\arcsec~pix$^{-1}$, and a field-of-view (FoV) of $9.1 \\times 9.9$\\arcsec\\ and $2.2 \\times 2.2$\\arcsec, for {\\em EMMI-R} and {\\em SuSI2}, respectively. Table \\ref{tab:ntt} summarises the imaging observations included in this article, lists the ESO filters used, and the total exposure times in each band. \\begin{table} \\caption{NTT observations} \\centering \\begin{tabular}{@{}llcc@{}} \\hline \\hline Observation & Camera & ESO filter \\# & ExpTime [ s ] \\\\ \\hline \\halpha & {\\em EMMI-R} & 598 & 900 \\\\ \\halpha\\ cont. & {\\em EMMI-R} & 597 & 1200 \\\\ \\hbeta & {\\em SuSI2} & 549 & 2866 \\\\ \\hbeta\\ cont. & {\\em EMMI-R} & 770 & 1800 \\\\ \\hline \\end{tabular} \\label{tab:ntt} \\end{table} Data were first reduced by the standard routines in {\\tt NOAO/IRAF}: bias subtraction and flat-field correction using well exposed sky- and dome-flats. Images were aligned using the {\\tt GEOMAP/GEOTRAN} tasks and smoothed to the seeing of the worst seeing image using the {\\tt GAUSS} task. All {\\em HST} bandpasses are aligned with the {\\em NTT} frames and the PSF is degraded by convolution with a Gaussian kernel. \\halpha\\ and \\hbeta\\ were continuum subtracted to obtain the nebular emission fluxes, accounting for contamination by [\\nii] (\\halpha\\ only) and underlying stellar absorption. [\\nii] contamination is estimated using [\\nii]$\\lambda 6583$\\AA/\\halpha$=0.189$ \\citep{bergvall02}. Stellar absorption is estimated from the best-fitting {\\em Starburst99} spectrum at each pixel by modifying the spectral fitting code described in Sect.~\\ref{sect:contsub}. The {\\em Starburst99} stellar libraries \\citep{mart05} were built upon the latest model atmospheres and include full line-blanketing for all stars and non-LTE effects for hot stars, thereby providing the best estimate of the stellar absorption features available. With the age of the stellar population we measure the equivalent width of \\hbeta\\ using the same line and continuum wavelength windows as the Lick index \\citep{lick}. The windows used for \\halpha\\ are of the same size as those for \\hbeta, but scaled to \\halpha\\ (i.e. the same windows $\\times 6563/4861=1.35$). \\ebv\\ is generated from \\halpha/\\hbeta, assuming a temperature of 10,000~K and intrinsic line ratio of 2.86, using the extinction law of the SMC following \\cite{Gordon03,Fitzpatrick_Massa88}. \\subsection{X-Ray: {\\em Chandra} and {\\em XMM-Newton}}\\label{sect:resxray} Haro\\,11 happens to be located 14\\arcmin\\ away from the Cartwheel galaxy, of which X-ray observations have been obtained with both the {\\em Chandra} and {\\em XMM-Newton} telescopes \\citep{wolt04}. Respective total integration times in {\\em Chandra} and {\\em XMM-Newton} were 80 and 70 ksec. Regarding {\\em XMM-Newton}, Haro\\,11 only falls within the field--of--view of the {\\em MOS2} detector, having unfortunately been missed by both the {\\em MOS1} and {\\em PN} chips. Both the {\\em Chandra} and {\\em XMM-Newton} datasets have been obtained from their respective archives. Since both telescopes operate with curved focal-plane configuration, the point spread function degrades severely with angular distance from the optical axis. For the {\\em XMM-Newton} detection, the PSF is so distorted that all spatial information has been lost, but the object is bright enough that a one-dimensional spectrum can be extracted from the {\\em MOS2} data. In the {\\em Chandra} observation, Haro\\,11 lies on the S1 chip: the object is clearly detected and we were able to perform a spatial analysis of the source (see Sect.~\\ref{sect:res_xray}). The {\\em Chandra} data were processed following the X-Ray Data Centre pipeline software. The level 1 events were reprocessed using {\\tt Ciao~3.4} task {\\tt acis\\_process\\_events}. The whole band (0.2-10 keV) image was smoothed applying the {\\tt csmooth} task. The image was smoothed with a minimal signal--to--noise of 3 using a circular Gaussian kernel. The source spectrum was extracted from a circular region of radius 30\\arcsec. The background was obtained from the combination of four circular regions located close to the source and avoiding other X-ray sources in the field. Source counts were grouped to have at least 20 counts per bin to allow modified $\\chi^2$ minimisation technique \\citep{kendall73} in the spectral analysis. Redistribution matrix and auxiliary response matrix files were generated. The {\\em XMM-Newton} data were processed using the standard {\\tt Science Analysis System, SAS, v.7.0.0} \\citep{gabriel04}. The most up-to-date files available as of January 2007 were used for the reduction process. The time intervals corresponding to high background events were removed using the method followed in \\cite{piconcelli04}. The resulting exposure for the {\\em MOS2} data is only of 44 ks. No sign of pile-up was found in the {\\em MOS2} image, according to the {\\tt epaplot SAS} task. A visual inspection of the 0.2-10~keV image shows a distorted shape due to location of the source, close to the edge of the CCD. Therefore, no further analysis of the {\\em XMM-Newton} image has been performed. The source spectrum was obtained, as for the {\\em Chandra} observation, from a circular region with a radius of 30\\arcsec. The background was extracted from a circular region close to the source and avoiding other X-ray sources in the field. Source counts were grouped to have at least 20 counts per bin. {\\tt SAS} appropriate tasks were used to generate the distribution matrix and auxiliary response matrix files. On-axis observations of Haro\\,11 have also been obtained with the {\\em Chandra} telescope \\citep{grimes07} although their archive status is still proprietary. Quantities derived from that dataset are, however, compatible with those presented here using off-axis archival observations (see Section~\\ref{sect:res_xray}). ", "conclusions": "\\label{sect:conc} Using {\\em HST/ACS} we have mapped, calibrated, and analysed the \\lya\\ emission from nearby luminous blue compact galaxy Haro\\,11. The \\lya\\ emission has been compared to: \\halpha; \\ebv\\ as determined through both \\halpha/\\hbeta\\ and UV continuum; the kinematic structure of the galaxy from previous studies; and the properties of the super star clusters. We have used archival X-ray data to map the hot outflowing gas and to put constraints on the evolutionary state of the burst. From SED fitting we have estimated and mapped the predicted Lyman-continuum production at 900\\AA, and compared this direct measurements. Most notably: \\begin{itemize} \\item{Our photometry reproduces spectroscopic fluxes as determined from the {\\em IUE} satellite. The total \\lya\\ flux is found to be $79.6 \\times 10^{-14}$~erg~s$^{-1}$~cm$^{-2}$, corresponding to a luminosity of $7.22 \\times 10^{41}$~erg~s$^{-1}$. } \\item{The escaping fraction of \\lya\\ photons is found to be $\\sim 3$\\%.} \\item{\\lya\\ shows almost no spatial correlation with \\halpha. The \\lya\\ morphology shows a central bright emission region where \\lya\\ photons escape the galaxy directly, surrounded by a low surface brightness diffuse halo that results from multiple resonant scatterings. 90\\% of the \\lya\\ output is in the halo component. \\lya/\\halpha\\ at the values predicted by recombination theory are seen only in the brightest central regions.} \\item{Little correlation is seen between \\lya\\ and dust. The main central \\lya\\ emitting region shows bright emission from a region where \\ebv\\ is significantly greater than regions that show only \\lya\\ absorption. In the regions of halo emission, \\ebv\\ extends beyond 1 although \\lya\\ is resonantly scattered and cannot feel the effect of such high extinction. This again indicates that dust is not the major regulatory factor governing the \\lya\\ morphology, which appears to be driven more by the H{\\sc i} distribution and its kinematic structure. } \\item{X-ray observations reveal a diffuse, soft component centred on the central star forming knots, and showing hot wind-driven gas pushed outside the central starburst region as probed by UV bands. The brightest X-ray regions are found to be spatially coincident with the regions of highest surface brightness in \\lya. This indicates that \\lya\\ emission from central regions may be the result of a perturbed and outflowing ISM, accelerated by the release of mechanical energy from the SSCs located around knot \\C. } \\item{From the super star clusters, we find that peak \\lya\\ equivalent widths are small at the youngest ages (1~Myr), reach a maximum from clusters in the range 2.5--4~Myr, and decline again to zero at ages $\\gtrsim8$~Myr. This is qualitatively consistent with models of \\lya\\ emission resulting from outflows in the neutral ISM. } \\item{Haro\\,11 is the only galaxy known to emit Lyman-continuum radiation ($\\lambda < 912$\\AA) in the nearby universe. From fitting spectral synthesis models we estimate the continuum flux at 900\\AA\\ to be $f_{900,0}=12.3\\times10^{-14}$~erg~s$^{-1}$~cm$^{-2}$~\\AA$^{-1}$. This corresponds to an escape fraction of 9\\% at 900\\AA. } \\end{itemize}" }, "0710/0710.1767_arXiv.txt": { "abstract": "{The shell-type supernova remnant RX J1713.7--3946 was observed during three years with the H.E.S.S. Cherenkov telescope system. The first observation campaign in 2003 yielded the first-ever resolved TeV gamma-ray image. Follow-up observations in 2004 and 2005 revealed the very-high-energy gamma-ray morphology with unprecedented precision and enabled spatially resolved spectral studies. Combining the data of three years, we obtain significantly increased statistics and energy coverage of the gamma-ray spectrum as compared to earlier H.E.S.S. results. We present the analysis of the data of different years separately for comparison and demonstrate that the telescope system operates stably over the course of three years. When combining the data sets, a gamma-ray image is obtained with a superb angular resolution of 0.06 degrees. The combined spectrum extends over three orders of magnitude, with significant gamma-ray emission approaching 100 TeV. For realistic scenarios of very-high-energy gamma-ray production, the measured gamma-ray energies imply efficient particle acceleration of primary particles, electrons or protons, to energies exceeding 100 TeV in the shell of RX J1713.7--3946.} \\begin{document} ", "introduction": "\\vspace{-0.4cm} \\begin{figure*} \\begin{center} \\includegraphics [width=0.78\\textwidth]{icrc0524_fig1.eps} \\end{center} \\caption{\\underline{\\textbf{Upper panel:}} H.E.S.S.\\ gamma-ray excess images from the region around RX~J1713.7$-$3946\\ are shown for three years. \\underline{\\textbf{Lower panel:}} 1D distributions generated from the non-smoothed, acceptance-corrected gamma-ray excess images. } \\label{fig1} \\end{figure*} The energy spectrum of cosmic rays measured at Earth exhibits a power-law dependence over a broad energy range. Starting at a few GeV $(1~\\mathrm{GeV} = 10^9~\\mathrm{eV})$ it continues to energies of at least $10^{20}~\\mathrm{eV}$. The power-law index of the spectrum changes at two characteristics energies: in the region around $3 \\times 10^{15}~\\mathrm{eV}$ -- the \\emph{knee} region -- the spectrum steepens, and at energies beyond $10^{18}~\\mathrm{eV}$ it hardens again. Up to the knee, cosmic rays are believed to be of Galactic origin, accelerated in shell-type supernova remnants (SNRs) -- expanding shock waves initiated by supernova explosions~\\cite{HillasReview}. However, the experimental confirmation of an SNR origin of Galactic cosmic rays is difficult due to the propagation effects of charged particles in the interstellar medium. The most promising way of proving the existence of high-energy particles in SNR shells is the detection of very-high-energy (VHE) gamma rays ($E > 100~\\mathrm{GeV}$), produced in interactions of cosmic rays close to their acceleration site~\\cite{DAV}. Recently the VHE gamma-ray instrument H.E.S.S. has detected two shell-type SNRs, RX~J1713.7$-$3946~\\cite{Hess1713a,Hess1713b} and RX~J0852.0--4622~\\cite{HessVelaJr_a,HessVelaJr_b}. The two objects show an extended morphology and exhibit a shell structure, as expected from the notion of particle acceleration in the expanding shock fronts. While it is difficult to attribute the measured VHE gamma rays unambiguously to nucleonic cosmic rays (rather than to cosmic electrons), the measured spectral shapes favour indeed in both cases a nucleonic cosmic-ray origin of the gamma rays~\\cite{Hess1713b,HessVelaJr_b}. Apart from the first unambiguous proof of multi-TeV particle acceleration in SNRs, the question of the highest observed energies remains an important one. Only the detection of gamma rays with energies of 100~TeV provides experimental proof of acceleration of primary particles, protons or electrons, to the \\emph{knee} region (1~PeV). Here we present a combined analysis of H.E.S.S.\\ data of RX~J1713.7$-$3946\\ of three years, from 2003 to 2005. A comparison of the three data sets demonstrates the expected steady emission of the source as well as the stability of the system. Special emphasis is then devoted to the high-energy end of the combined spectrum. ", "conclusions": "\\vspace{-0.4cm} The complete H.E.S.S.\\ data set of the SNR RX~J1713.7$-$3946\\ recorded from 2003 to 2005 is presented here. Very good agreement is found for both the gamma-ray morphology and the differential energy spectra over the years. The combined analysis confirms the earlier findings nicely: the gamma-ray image reveals a thick, almost circular shell with significant brightness variations. The spectrum follows a hard power law with significant deviations at higher energies (beyond a few TeV). In the combined image using $\\sim63$~hours of H.E.S.S.\\ observations an unprecedented angular resolution of $0.06\\degr$ is achieved. The high-energy end of the combined spectrum approaches 100~TeV with significant emission $(4.8\\sigma)$ beyond 30~TeV. Given the systematic uncertainties in the spectral determination at these highest energies and comparable statistical uncertainties despite the long exposure time, this measurement is presumably close to what can be studied with the current generation of imaging atmospheric Cherenkov telescopes. From the largest measured gamma-ray energies one can estimate the corresponding energy of the primary particles. In case of $\\pi^0$-decay gamma rays, energies of 30~TeV imply that primary protons are accelerated to $30~\\mathrm{TeV} / 0.15 = 200~\\mathrm{TeV}$ in the shell of RX~J1713.7$-$3946. On the other hand, if the gamma rays are due to Inverse Compton scattering of VHE electrons, the electron energies at the current epoch can be estimated in the Thompson regime as $E_\\mathrm{e}~\\approx~20~\\sqrt{E_\\gamma}~\\mathrm{TeV} \\approx 110~\\mathrm{TeV}$. At these large energies Klein--Nishina effects start to be important and reduce the maximum energy slightly such that $\\sim100~\\mathrm{TeV}$ is a realistic estimate. RX~J1713.7$-$3946\\ remains an exceptional SNR in respect of its VHE gamma-ray observability, being at present the remnant with the widest possible coverage along the electromagnetic spectrum. The H.E.S.S.\\ measurement of significant gamma-ray emission beyond 30~TeV without indication of a termination of the high-energy spectrum provides proof of particle acceleration in the shell of RX~J1713.7$-$3946\\ beyond $10^{14}$~eV, up to energies which start to approach the region of the cosmic-ray \\emph{knee}." }, "0710/0710.4468_arXiv.txt": { "abstract": "To study the effect of metallicity on the mass-loss rate of asymptotic giant branch (AGB) stars, we have conducted mid-infrared photometric measurements of such stars in the Sagittarius (Sgr dSph) and Fornax dwarf spheroidal galaxies with the 10-$\\mu$m camera VISIR at the VLT. We derive mass-loss rates for 29 AGB stars in Sgr dSph and 2 in Fornax. The dust mass-loss rates are estimated from the $K-[9]$ and $K-[11]$ colours. Radiative transfer models are used to check the consistency of the method. Published IRAS and Spitzer data confirm that the same tight correlation between $K-[12]$ colour and dust mass-loss rates is observed for AGB stars from galaxies with different metallicities, i.e. the Galaxy, the LMC and the SMC. The derived dust mass-loss rates are in the range 5$\\times10^{-10}$ to 3$\\times10^{-8}$ M$_{\\odot}$yr$^{-1}$ for the observed AGB stars in Sgr dSph and around 5$\\times10^{-9}$ M$_{\\odot}$yr$^{-1}$ for those in Fornax; while values obtained with the two different methods are of the same order of magnitude. The mass-loss rates for these stars are higher than the nuclear burning rates, so they will terminate their AGB phase by the depletion of their stellar mantles before their core can grow significantly. Some observed stars have lower mass-loss rates than the minimum value predicted by theoretical models. ", "introduction": "Stars with initial mass in the range 0.8--8 M$_{\\odot}$ go through an asymptotic giant branch (AGB) phase towards the end of their evolution. This evolutionary phase is dominated by strong mass loss. The star expels material at rates up to $10^{-4}M_{\\odot}$yr$^{-1}$, eventually ejecting between 20 and 80 per cent of its initial main-sequence mass. This process is of great importance for the chemical evolution of our Galaxy. Mass loss from AGB stars contributes to around half of the gas recycled by stars (Maeder 1992), and is a dominant source of Galactic dust. The mass-loss mechanism of AGB stars is a two-step process. First, shocks due to pulsations from the star produce gas over-densities in an extended atmosphere (e.g. H\\\"ofner et al. 1998). This triggers the formation of dust. Secondly, radiation pressure accelerates the dust to the escape velocity. The gas is carried along through friction with the dust particles. Pulsations alone can explain mass-loss rates up to about $10^{-7}M_{\\odot}$yr$^{-1}$, but the much higher rates observed require dust-driven winds (Bowen \\&\\ Wilson 1991). The importance of metallicity on the mass-loss rates of AGB stars is not well understood. In low metallicity environments less dust is expected to form, leading to lower predicted mass-loss rates. Theoretical work by Bowen \\& Willson (1991) predicts that for metallicities below [Fe/H]$=-1$ dust-driven winds fail, and the wind becomes pulsation-driven. This would affect the evolution of a low metallicity host galaxy in two obvious ways. First the stellar dust production would be much reduced and therefore the composition of the material out of which new stars and planets were forming would be significantly different. Secondly the much weaker dust-driven winds allow the degenerate core of the AGB star to grow for longer, resulting in much higher masses for the remnant white dwarfs. In extreme cases, the core could reach the Chandrasekhar limit before mass loss terminates its evolution. The Bowen \\&\\ Willson mass-loss rates predict the occurrence of AGB supernovae at very low metallicities (Zijlstra 2004). It is therefore important to test whether dust-driven winds exist at low metallicity. Observations in the Magellanic Clouds and the Galaxy have shown that the dust mass-loss rates are smaller in the Magellanic Clouds. Assuming that the dust-to-gas ratio is a linear function of metallicity ([Fe/H]=$-0.6$ (Venn 1999) and [Fe/H]=$-0.3$ for the Small and Large Magellanic Clouds respectively) yields the conclusion that the total mass-loss rate (dust+gas) may be similar in the three galaxies (van Loon 2000, 2006), although this assumption may not be strictly correct, as the dust in the carbon stars comes from primary carbon. In order to obtain further constraints on the effect of metallicity on the mass-loss rates from AGB stars, and to know if dust-driven mass loss can occur at very low metallicities, we need to study the mass-loss rates from more metal-poor AGB stars. The dusty circumstellar envelopes surrounding these AGB stars absorb the radiation from the central star and re-emit it in the thermal infrared. Observations at infrared wavelengths have enabled the study of mass-loss from Galactic and Magellanic Clouds AGB stars. Thanks to the sensitivity achievable with mid-infrared instruments on 8m class ground-based telescopes, it is now possible to perform such studies for AGB stars in more distant Local Group galaxies. We thus observed AGB stars in the Fornax dSph (Fornax) and Sagittarius (Sgr dSph) dwarf spheroidal galaxies using the mid-infrared camera VISIR on the VLT (ESO, Chile). Fornax is a satellite galaxy of the Milky Way, at a distance of $\\sim$ 138 kpc. Sgr dSph is located behind the Galactic disc and bulge at a distance of $\\sim$24 kpc. It is currently being torn apart by tidal forces (Majewski et al. 2003). Those galaxies have low metallicities (see section 2). Furthermore the distances of both galaxies are quite well known, making an estimation of the mass-loss rates easier than it is for stars within the Galaxy. ", "conclusions": "We have presented a study of mass-loss from AGB stars in the Sagittarius and Fornax dwarf spheroidal galaxies using mid-infrared photometry obtained with VISIR (VLT, ESO). We have shown that the well known relation between $K-[12]$ colours and mass-loss rates among Galactic stars is also valid for AGB stars in lower metallicity galaxies. We used this relation and radiative transfer models to estimate dust mass-loss rates for 15 stars in Sgr dSph and 2 stars in Fornax. Our study showed that some AGB stars in the Sgr dSph and Fornax galaxies are losing mass. The estimated dust mass-loss rates are in the range 5$\\times10^{-10}$ to $3\\times10^{-8}$ M$_{\\odot}$yr$^{-1}$ for the stars in Sgr dSph and around 5$\\times10^{-9}$ M$_{\\odot}$yr$^{-1}$ for those in Fornax. The values obtained with the two different methods are of the same order of magnitude. These mass-loss rates are higher than the nuclear burning rates, so these stars will terminate their AGB evolution by the depletion of their stellar mantles, before their cores can grow significantly. Using previous work to get an estimation of the dust-to-gas mass ratio for the stars observed we found that most of them had very low mass-loss rates for their luminosity. This is in contradiction with theoretical predictions (see e.g Gail \\& Sedlmayr 1987)." }, "0710/0710.4142_arXiv.txt": { "abstract": "We infer the large-scale source parameters of dusty galaxies from their observed spectral energy distributions (SEDs) using the analytic radiative transfer methodology presented in Chakrabarti \\& McKee (2005). For local ultra-luminous infrared galaxies (ULIRGs), we show that the millimeter to far-infrared (FIR) SEDs can be well fit using the standard dust opacity index of 2 when self-consistent radiative transfer solutions are employed, indicating that the cold dust in local ULIRGs can be described by a single grain model. We develop a method for determining photometric redshifts of ULIRGs and sub-mm galaxies from the millimeter-FIR SED; the resulting value of $1+z$ is typically accurate to about $10\\%$. As such, it is comparable to the accuracy of near-IR photometric redshifts and provides a complementary means of deriving redshifts from far-IR data, such as that from the upcoming $\\it{Herschel~Space~Observatory}$. Since our analytic radiative transfer solution is developed for homogeneous, spherically symmetric, centrally heated, dusty sources, it is relevant for infrared bright galaxies that are primarily powered by compact sources of luminosity that are embedded in a dusty envelope. We discuss how deviations from spherical symmetry may affect the applicability of our solution, and we contrast our self-consistent analytic solution with standard approximations to demonstrate the main differences. ", "introduction": "The far-infrared (FIR) spectral energy distribution (SED) is a vital implement in understanding the physical conditions of dusty sources. Chakrabarti \\& McKee (2005, henceforth CM05), presented self-consistent analytic radiative transfer solutions for the spectra of unresolved, homogeneous, spherically symmetric, centrally heated, dusty sources. We showed that from two colors in the millimeter (mm) and FIR portion of the spectrum one can approximately infer the mm to FIR, and that this in turn determines the luminosity to mass ratio, $L/M$, and surface density, $\\Sigma$, which (at low redshift) are distance-independent parameters. With a distance measurement, one can further infer the size, mass, and luminosity of the source. We extensively compared our analytic solutions against a well-tested numerical scheme, DUSTY (Ivezic \\& Elitzur 1997), to find excellent agreement with the numerical results. Here, we apply this methodology to dusty galaxies to derive their large-scale source parameters. We discuss applications to protostars and radiative transfer methodology of clumpy envelopes in a separate forthcoming paper. From observations of their SEDs, the IRAS all-sky survey characterized ULIRGs as a class of extremely luminous ($L_{8-1000~\\micron}>10^{12}L_{\\odot}$) galaxies that emit most of their energy in the FIR (Soifer et al. 1984; Aaronson \\& Olszewski 1984; Soifer et al. 1987; Sanders \\& Mirabel 1996). These galaxies were then understood to be a new class of objects, quite distinct from those studied by optical surveys as little correlation was found between their optical and infrared luminosities (Sanders \\& Mirabel 1996). That the demographics of these galaxies is not simply an extrapolation of normal galaxies can be seen from the fact that the luminosity function on the bright end ($L_{\\rm IR} \\ga 1 \\times 10^{11} L_{\\odot}$) is significantly in excess of the Schechter function (Rieke \\& Lebofsky 1986). Theoretical models have suggested that heavily starbursting systems like ULIRGs can be produced via mergers of roughly equal mass galaxies (Toomre \\& Toomre 1972; Mihos \\& Hernquist 1996), and recent observations corroborate the idea that local ULIRGs are products of major mergers (Dasyra et al. 2006). While the limited sensitivity of IRAS did not allow for a characterization of the redshift evolution of this dusty, luminous population of galaxies, observations by the $\\it{Spitzer~Space~Telescope}$ indicate that there is a strong evolution in this population (on the basis of near and mid-IR observations converted to total IR luminosities using observational templates) out to $z \\sim 1$ (e.g. Le Floch et al. 2005), with the contribution from ULIRGs to the comoving IR energy density increasing by an order of magnitude from local systems to $z \\sim 1$. Understanding the FIR SEDs of dusty galaxies has renewed importance today. Submillimeter galaxies (SMGs; $F_{850~\\micron} \\ga 1~\\rm mJy$) (Ivison et al. 2000; Blain et al. 1998) are luminous ($L \\ga 10^{12} L_{\\odot}$), dusty galaxies at moderate redshifts (the median redshift in the Chapman et al. 2003, 2005 samples is $z \\sim 2$). They are faint at optical wavelengths and were discovered in the first deep extragalactic surveys in the sub-mm wavebands (the SCUBA Cluster Lens Survey; Smail et al 1997, 2002). SMGs produce a significant fraction of the energy output of the high redshift early universe, and hence represent a cosmologically significant population (Smail et al 1997, Blain et al 2002, Blain et al 1999). Chapman et al. (2005) find that SMGs and Lyman break galaxies contribute equally to the star formation density at $z \\sim 2-3$ and that, when extrapolated to lower fluxes, SMGs may be the dominant site of massive star formation at this epoch. Upcoming instruments, such as the $\\it{Herschel~Space~Observatory}$ and SCUBA-2, will be able to perform routine observations of SMGs at rest-frame FIR wavelengths, which is critical for observationally determining the bolometric luminosities of this high redshift, cosmologically significant population. In contrast to local ULIRGs, the suggested formation mechanisms and evolutionary scenarios for SMGs remain varied in nature, ranging from primeval, heavily accreting galaxies undergoing a starburst (Rowan-Robinson 2000; Efstathiou \\& Rowan-Robinson 2003) to products of gas-rich major mergers undergoing intense feedback (Chakrabarti et al. 2007b), with recent observations favoring the latter scenario (Nesvadba et al. 2007; Bouche et al. 2007). In particular, Bouche et al. (2007) suggest that dissipative major mergers may have produced the SMG population on the basis of their finding that the SMG population has lower angular momenta and higher matter densities compared to the UV/optically selected population. A self-consistent analytic method of inferring source parameters from the observed SEDs may be an useful alternative to SED templates in analyzing upcoming FIR data sets. We give the general relations for observed quantities in terms of the redshift, and graphically depict the variation of these quantities with redshift, which is significant even at $z\\sim 1$. We show that this implies that one can estimate the value of $1+z$ for ULIRGs and SMGs from the mm and FIR SED with an accuracy $\\sim 10\\%$ (this is comparable to typical accuracies of photometric redshift codes, which estimate $z$ to typical accuracies of $\\sim 10\\%$, e.g. the IMPZ code of Babbedge et al. 2004, or the widely used HYPERZ code of Bolzonella et al. 2000), given that our assumptions are satisfied: (1) the mm-FIR spectrum is due to reprocessing of emission from a central, dust-enshrouded source; (2) the dust can be approximated as being homogeneous and spherically distributed, with a density that varies as a power of the radius; (3) the source is sufficiently opaque that emission from the dust destruction front is negligible; and (4) the luminosity-to-mass ratio, $L/M$, of high-redshift ULIRGs and SMGs is similar to that of low-redshift ULIRGs (this last assumption is verified for the small set of SMGs for which adequate data currently exist). Chapman et al. (2005) point out that the dust temperatures inferred for SMGs are significantly lower than those of local ULIRGs and conclude that FIR photometric redshifts have an uncertainty $\\Delta z \\simeq 1$; as we shall see, our method is significantly more accurate. We primarily focus our discussion of SMGs on sources observed recently by Kovacs et al. (2006) using the $350~\\micron$ band of SHARC-2, which is currently the most direct observational probe of the rest-frame FIR of high redshift SMGs. The organization of the paper is as follows: in \\S 2, we review the basics of the analytic radiative transfer methodology presented in CM05 and collect the main expressions in Appendix A; in \\S 2.1, we explain the general procedure of applying our results, and contrast our solution with standard approximations in Appendix B. \\S 3 is devoted to a treatment of ULIRGs, where we infer the large scale parameters of a dozen local ULIRGs by fitting to the FIR SED. In \\S 4 we present SED fits for a sample of SMGs. In \\S 5 we present the principal result of this paper, a method of inferring redshifts from FIR SEDs, and we demonstrate its applicability both with a simulated test case and with data of SMGs. We conclude in \\S 6. In Appendix B we discuss standard approximations, such as the Hildebrand (1983) prescription for the mass in terms of the mm flux, and modified blackbody single temperature models in fitting the SEDs of ULIRGs and protostars (Yun \\& Carrilli 2002, henceforth YC02). For purposes of illustration, we graphically contrast our solution and the standard approximations against the numerical results from DUSTY over the astrophysical parameter space. ", "conclusions": "We have applied the shell methodology for radiative transfer developed in CM05 to a range of extragalactic sources, from local ULIRGs to high redshift SMGs. The main results are: 1. Using the general expressions for the SEDs of dusty sources given in CM05, we have shown how to derive the light-to-mass ratio, $L/M\\delta$, and the mean surface density, $\\Sigma\\delta$, of dusty galaxies at cosmological distances. Here $M$ and $\\Sigma$ refer to the gas mass, and $L$ is the FIR luminosity as determined from observations at rest wavelengths $\\ga 60$~\\micron. The effective dust-to-gas ratio, $\\delta$, is the ratio of the actual FIR opacity to the one we have adopted, which is twice the Weingartner \\& Draine (2001) opacity; the factor 2 allows for ice mantles. Approximate expressions are given for the case in which the radius of the dusty envelope, $R_c$, is much larger than the characteristic photospheric radius, $\\rch$ (i.e., for $\\rct\\equiv R_c/\\rch\\gg 1$). 2. The long-wavelength slope of ULIRGs can be fit with a standard dust opacity curve ($\\kappa_\\nu\\propto \\nu^2$ in the FIR) when a self-consistent radiative transfer solution is employed. This confirms the conclusion of Dunne \\& Eales (2001), who used two-temperature fits to the SED. Comparison with the three sources in DS98 for which there are resolved measurements for the galaxies in our sample (Arp 220, Mrk 231 and IRAS 10565+2448) shows that on average our mass estimates are about 1.3 times greater than theirs, which is excellent agreement in view of the approximations in each method and the possible variations in the value of $\\delta$ from galaxy to galaxy. 3. From an analysis of the FIR data of 10 local ULIRGs, we find that the mm to FIR SED can be well-described by our simple construct of a spherically symmetric dust envelope surrounding a central source of luminosity. We find that a density profile $\\rho \\propto r^{-2}$ provides the best to the FIR data and is consistent with CO masses. We report our findings for the luminosities, masses, and sizes of these 10 ULIRGs. 4. We find that local ULIRGs and high redshift SMGs ($z \\sim 2$) have similar $L/M\\delta$ ratios, with the SMGs in our sample having values $\\sim 1.45$ times smaller. The SMGs in our sample have luminosities about 5.5 times larger, and masses about 8 times larger, than the ULIRGs in our sample. 5. We have developed a method of inferring FIR photometric redshifts. The accuracy of the method depends on whether the galaxy has a light-to-mass ratio comparable to the template, but since $1+z$ scales as only the 1/6 power of $L/M\\delta$, significant variations are allowed. Using Arp 220 as an example, we showed that our method should work for redshifts $z\\la 10$. We tested our method on a sample of five SMGs for which good photometric data are available. Under the assumption that the SMGs had the same light-to-mass ratio as a sample of local ULIRGs, we were able to infer values of $1+z$ for the SMGs with an average accuracy of 10\\% and a worst accuracy of 15\\%. If we used the average light-to-mass ratio for the SMGs, the average accuracy improved to about 5\\%. As the samples of high-redshift galaxies grow, our knowledge of both the typical light-to-mass ratio and the accuracy of FIR photometric redshifts will improve. Our method should be applied to galaxies that radiate most of their energy in the FIR, such as ULIRGs and SMGs. Whether this condition is met can be determined by confirming that the near-IR (or optical/UV) luminosity is significantly less than that emerging in the FIR. Our method works even if AGN provide a significant fraction of the luminosity, although our sample is not large enough to determine how large the AGN fraction can become before our method breaks down. As our method is analytic, it can be employed to quickly obtain photometric redshifts of large samples of SMGs, as are expected to be detected in the FIR by the upcoming $\\it{Herschel}$ mission. This FIR photometric redshift method provides a complementary means of inferring the redshift when near-IR methods are not available or are not viable. 6. We discuss how our approximation for the radiative transfer compares with the standard single-temperature SED in Appendix B. The accuracy of the single-temperature blackbody approximation degrades for extended envelopes, $\\rct \\ga 100$, but the approximation is typically accurate to within a factor 2 for compact envelopes, for which the source function does not probe a large range of temperatures. The effective dust temperature in the single-temperature approximation is close to the outer core temperature. \\bigskip \\bigskip" }, "0710/0710.1677_arXiv.txt": { "abstract": "We report the negative results from a search for 6.7 GHz methanol masers in the nearby spiral galaxy M33. We observed 14 GMCs in the central 4 kpc of the Galaxy, and found 3$\\sigma$ upper limits to the flux density of $\\sim$9 mJy in spectral channels having a velocity width of 0.069 \\kms. By velocity shifting and combining the spectra from the positions observed, we obtain an effective 3$\\sigma$ upper limit on the average emission of $\\sim$1mJy in a 0.25 \\kms\\ channel. These limits lie significantly below what we would expect based on our estimates of the methanol maser luminosity function in the Milky Way. The most likely explanation for the absence of detectable methanol masers appears to be the metallicity of M33, which is modestly less than that of the Milky Way. ", "introduction": "6.7 GHz methanol masers are the second brightest maser transition ever observed, and are typically much brighter than OH masers. The properties of methanol masers in the Magellanic clouds \\citep{sinclair1992,ellingsen1994b,beasley1996} are consistent with those of our Galaxy, given appropriate consideration for different galactic properties such as metallicity. The SMC has an oxygen abundance 12 + log(O/H) = 7.96 \\citep{vermeij2002}, which is a factor $\\simeq$ 6 smaller than that of GMCs in the Milky Way \\citep{peimbert1993}. However, 6.7 GHz methanol masers have not been discovered in galaxies beyond the Magellanic clouds. In particular, surveys have shown that there is no analog to OH megamasers at 6.7 GHz \\citep{ellingsen1994a,phillips1998,darling2003}. It would be of interest to detect methanol masers in a Milky Way like galaxy. If the number of sources detected were large, one could derive the methanol maser luminosity function since all sources would be at the same distance. Further, one could look for correlations of methanol masers with giant molecular cloud masses, other types of masers, etc. The Arecibo radio telescope offers unequaled sensitivity for targeted surveys for methanol masers, and the nearest spiral galaxy which can be observed with this instrument is M33. It is difficult to develop an optimal strategy to search for extragalactic methanol masers since the luminosity function of methanol masers in our Galaxy is unknown. This is mostly due to difficulties in determining distances to methanol masers, which is compounded by the kinematic distance ambiguity. We have adopted the following approach. We selected sources from the general catalog \\citep{pestalozzi2005}, eliminating sources wihtin 10 degrees of longitudes 0 and 180 degrees since these suffer from large uncertainties in kinematic distance. Assuming that all masers are at their near kinematic distance, we then scaled the source flux density to the distance of M33, which we take to be 840 kpc, averaging the results of \\cite{lee2002} and \\cite{freedman1991}, which gives a result consistent with the distance determined by \\cite{kim2002}. The resulting histogram of ``expected'' methanol maser flux densities is showin in Figure \\ref{expected_hist} \\begin{figure} \\begin{center} \\includegraphics[angle=0]{f1.eps} \\caption{Distribution of expected flux densities of methanol masers in M33 based on observed masers in the Milky Way.} \\label{expected_hist} \\end{center} \\end{figure} We thereby obtained an estimate for the distribution of flux density of methanol masers in M33, which of course must be regarded with considerable skepticism, as the original Galactic sample has significant biases, not to mention the unknown differences in the properties of massive star formation in M33 and the Milky Way. It is also inherently pessimisitic in the sense that some of the Galactic masers are at their far kinematic distance, and hence are more luminous than we have assumed. With these caveats, we found a ``high flux density tail'' for our hypothetical methanol maser population in M33 that extends from 100 mJy down to 1 mJy. Approximately 16\\% of the total number of masers would be expected to have flux density greater than 1 mJy, with about 6\\% having flux density greater than 3 mJy. ", "conclusions": "It is difficult to be highly quantiative about the lack of detection of methanol maser emission in M33 beyond the statements given above. Given that the 40\\arcsec\\ Arecibo beam subtends a region 160 pc in size at the distance of M33, we should be sensitive to a methanol maser located anywhere within the individual giant molecular clouds being observed. Based on our estimate of the Milky Way methanol maser luminosity function, we would expect a significant fraction of masers to have a flux density $\\geq$ 10 mJy, significantly higher than the individual limits we have obtained, and an order of magnitude greater than the averaged limit derived by combining the results from the 14 GMCs observed. Explanations of the rarity of methanol masers in external galaxies have focused on (1) insufficient methanol density over path where amplification could take place and (2) insufficient pumping to invert the methanol transition in question \\citep{phillips1998}. Both of these can result from low metallicity. A reduced abundance of oxygen, for example, will likely reduce both the abuance of methanol (CH$_3$OH) and of dust, which is required to convert the short wavelength radiation from massive young stars to the infrared wavelengths required for maser pumping. The rarity of masers in the Magellanic clouds has been discussed in similar terms by \\cite{beasley1996}. The O/H ratio in the Magellanic clouds is dramatically lower than that of the Milky Way (see discussion in Beasley et al. 1996 and the more recent measurement by Vermeij \\& van der Hulst 2002). For M33, the O/H ratio is only slightly less than that of the Milky Way. \\cite{vilchez1988} determined that 12 + log(O/H) = 9.0 at the center of M33, falling to $\\sim$8.5 at distances between 2 and 5 kpc from the center of the galaxy. The best fit line of \\cite{crockett2006} to their new data plus previous data has a very small slope of -0.12 dex kpc$^{-1}$ and 12 + log(O/H) = 8.3 at 5 kpc from the center of M33. These values may be compared to Galactic values of 12 + log(O/H) = 8.51 for Orion and 8.78 for M17 \\citep{peimbert1993}. The GMCs we have observed are located between 0.5 and 4 kpc from the center of M33, with a mean distance of 1.9 kpc. We conclude that the relevant O/H ratio in M33 is between 0.1 and 0.4 dex below that of the Milky Way. This difference is much smaller than that between the SMC and the Milky Way, for which log(O/H) differs by $\\simeq$ 0.8 dex. If the relatively small difference between the metallicity in the Milky wand M33 is responsible for the lack of methanol masers in the latter, it suggests that the maser luminosity must be a very senstive function of the galactic metallicity. \\cite{henkel1987} have detected thermal methanol emission from two galaxies, IC342 and NGC253, finding that the fractional abundance of methanol is $\\simeq$10$^{-7.5}$. This is similar to that found in GMCs in the Milky Way, but is a factor of at least 100 lower than that required for high brightness methanol maser luminosity as discussed by \\cite{sobolev1997}. However, this difference is also found in comparing hot cores in the Milky Way, presumed to be the sites of methanol masers, with more extended molecular cloud regions. The methanol abundance may be greatly increased in regions near hot stars by thermal desorption of molecules frozen onto grain surfaces. We have no direct evidence regarding the methanol abundance in M33, so it is possible that the lower O/H ratio does yield an insufficient methanol abundance to produce highly luminous masers. There are individual regions within M33, presumably powered by massive young stars \\citep{hinz2004}, which are powerful far--infrared sources. The infrared flux is a critical requirement for maser pumping as elaborated in models of \\cite{cragg1992}, \\cite{sobolev1994}, and \\cite{sobolev1997}. The transition to the second torsional excited state occurs at a wavelength of $\\simeq$ 30 \\mic, so that dust temperatures of at least 150 K are required to achieve high maser brightness \\citep{sobolev1997}. Among our targets was number 8 of \\cite{engargiola2003}, which lies within 15\\arcsec of the optical nebula NGC 604. It has an IRAS luminosity of 6.8$\\times$10$^7$ \\Ls \\citep{rice1990} and the associated GMC has a mass derived from CO of 4.4$\\times$10$^5$ \\Ms \\citep{engargiola2003}. By Galactic standards this region would seem likely to harbour a high luminosity methanol maser, but no such emisson was detected. \\cite{fix1985} were unsuccessful in a search using the VLA for highly luminous OH masers in M33. Their limit of $\\simeq$ 25 mJy (5$\\sigma$) was sufficient to elminate the presence of any type I maser as luminous as the brightest type I masers in the Milky Way The present work thus reemphasizes the mystery of the lack of luminous masers in M33." }, "0710/0710.5923_arXiv.txt": { "abstract": "The primary optical caustic surface behind a Kerr black hole is a four-cusped tube displaced from the line of sight. We derive the caustic surface in the nearly asymptotic region far from the black hole through a Taylor expansion of the lightlike geodesics up to and including fourth-order terms in $m/b$ and $a/b$, where $m$ is the black hole mass, $a$ the spin and $b$ the impact parameter. The corresponding critical locus in the observer's sky is elliptical and a point-like source inside the caustics will be imaged as an Einstein cross. With regard to lensing near critical points, a Kerr lens is analogous to a circular lens perturbed by a dipole and a quadrupole potential. The caustic structure of the supermassive black hole in the Galactic center could be probed by lensing of low mass X-ray binaries in the Galactic inner regions or by hot spots in the accretion disk. ", "introduction": "Black holes (BHs) are among the most fascinating predictions of the general theory of relativity. Recent progresses in mass measurements of compact objects have supplied compelling evidence for their existence. Nearly every galaxy is supposed to contain a very massive BH at its center in the mass range from $10^{6}$ to $10^{9}~M_\\odot$ \\citep{na+qu05}. Despite strong hints that massive BHs mostly rotate rapidly, their spins are still unmeasured so that a full characterization of their properties is not possible. Analyses of luminosities and spectra of accretion disks around spinning BH are very promising tool, but the interpretation can be sometimes unclear due to uncertainties in the physics of the inflowing gas. The classical test of gravitational deflection of light rays passing near compact bodies can provide an alternative probe. In fact, the theoretical bases of lensing are well understood and, on the observational side, the supermassive BH supposed to be hosted in the radio source Sgr~A* in the Galactic center, with a mass of $\\sim 3.6\\times 10^6 M_\\odot$ and at a distance of $7.6~\\mathrm{Kpc}$ from the Earth \\citep{eis+al05}, offers a very appealing target for future space- and ground-based experiments. Lensing by rotating BHs has been studied from the very beginning of the Kerr spacetime \\citep{bo+li67}. Angular momentum first appears in gravitational lensing through terms $\\sim m a /b^2$ in the deflection angle. Up to this order, light deflection is well understood. Very different approaches can be undertaken \\citep{pi+ro77,ep+sh80,iba83,ba+ho04,bra86,dym86,gli99,kop97,ko+sc99,ko+ma02,as+ka00,ser03b} to show the degeneracy between a Kerr BH and a suitably displaced Schwarzschild lens \\citep{asa+al03,we+pe07}. Such analyses can be easily extended to general spinning mass distributions \\citep{ser02,se+ca02,ser03,ser05,ser07} and show many interesting features in the magnification pattern and in the image properties \\citep{ser03,ser05,we+pe07} . Investigations up to higher orders require the full consideration of the lightlike geodesics \\citep{car68,cha83}. The optical structure of the primary caustic surface \\cite{ra+bl94} as well as the appearance of both stars \\citep{cu+ba73} and accretion disk \\citep{vie93,be+do05} orbiting a Kerr BH have been detailed through numerical investigations. A clear analytical picture of the relativistic caustics in the strong deflection limit has also emerged \\citep{boz+al05,bo+sc07}. The missing part, which we are going to provide here, is an analytical treatment of map singularities and caustics in the weak deflection limit. This is the prerequisite for the study of critical points and lensing map inversion near them. The paper is organized as follows. In Section~\\ref{sec:geod}, we recall basic properties of lightlike geodesics in Kerr spacetime. Sections~\\ref{sec:caus} discusses the singularities of lens mapping and lensing near caustics. Section~\\ref{sec:disc} is devoted to some final considerations. In this paper, we will use units $G=c=1$, with $c$ the light speed in the vacuum, so that the constant $m (\\equiv G M /c^2)$ is the gravitational radius. ", "conclusions": "\\label{sec:disc} The analytical derivation performed in this paper corroborates and extends knowledge about the primary caustics which, in a earlier work \\citep{ra+bl94}, have been investigated only numerically and provides the first study, to our knowledge, of critical points and lens mapping near them. Our approach is the natural complement to qualitative methods which give some information on lensing properties without actually solving the equation for lightlike geodesics, such as the Morse theory \\citep{ha+pe06}. Together with the theoretical motivation, an equally compelling reason for investigating gravitational lensing in a Kerr spacetime comes from lensing observations towards Sgr~A* which are coming into the reach of observability and for which the weak-deflection limit approximation at the lowest order is not applicable. The stars surrounding the Galactic center have been considered as suitable targets for detection of lensing effects. Sgr~A* is expected to be lensing nearly ten sources at any given time for observations down to $K \\sim 21$ \\citep{jar98,al+st99,cha+al01} . Considering the only few stars whose orbital parameters have already been accurately determined, detectable lensing events are expected to occur in a temporal span of $\\sim 30$~years \\citep{bo+ma05}. The typical radius of these sources ($\\sim 10^{8}-10^{9}~\\mathrm{m}$) is much larger than the caustic width, so that the effects of the finite astroid size are washed out. Nevertheless, the astrometric shift of the caustic center, together with the precession orbital effects related to the spinning central body, must be accounted for when considering future experiments. Other appealing sources for lensing by Sgr~A* are low mass X-ray binaries (LMXBs), whose emission mostly originates in a region a few tens of kilometers across, consisting of the inner accretion disk around the BH accreting from the companion. Tens of thousands of stellar-masses BHs and neutron stars are likely to have settled dynamically into the central parsec of the Galaxy and perhaps few hundreds of them might have stellar companions \\cite{mun+al05}. Several tens of LMXBs have been already detected in the very inner regions down to a minimum projected distance of only $0.1~\\mathrm{pc}$ \\citep{mun+al05b}, with a detected overabundance of transients X-ray binaries within $1~\\mathrm{pc}$ \\citep{mun+al05}. These sources have been considered for detection of relativistic images \\citep{boz+al05} and could as well, since their small radius, probe the primary caustic surface. Compact emission regions in the clumpy and unsteady accretion flow near Sgr~A* are other interesting lensing sources \\citep{br+lo05,br+lo06}. Relativistic images of orbiting bright-spots could be detected in the near future with observations achieving submilliarcsecond resolution at infrared and submillimetre wavelengths. However, we remark that in our approximation the source distance must be large and close inner orbits can not be considered. The finite size of the primary caustics behind Sgr~A* could then be probed by lensing of either LMXBs or compact hot spots in the accretion flow whereas the effects of angular momentum, such as the shift in the caustic position, affect even lensing of main sequence stars in the Galactic center cluster." }, "0710/0710.5112_arXiv.txt": { "abstract": "Physical sizes of extended radio galaxies can be employed as a cosmological ``standard ruler'', using a previously developed method. Eleven new radio galaxies are added to our previous sample of nineteen sources, forming a sample of thirty objects with redshifts between 0 and 1.8. This sample of radio galaxies are used to obtain the best fit cosmological parameters in a quintessence model in a spatially flat universe, a cosmological constant model that allows for non-zero space curvature, and a rolling scalar field model in a spatially flat universe. Results obtained with radio galaxies are compared with those obtained with different supernova samples, and with combined radio galaxy and supernova samples. Results obtained with different samples are consistent, suggesting that neither method is seriously affected by systematic errors. Best fit radio galaxy and supernovae model parameters determined in the different cosmological models are nearly identical, and are used to determine dimensionless coordinate distances to supernovae and radio galaxies, and distance moduli to the radio galaxies. The distance moduli to the radio galaxies can be combined with supernovae samples to increase the number of sources, particularly high-redshift sources, in the samples. The constraints obtained here with the combined radio galaxy plus supernovae data set in the rolling scalar field model are quite strong. The best fit parameter values suggest that $\\Omega_m$ is less than about 0.35, and the model parameter $\\alpha$ is close to zero; that is, a cosmological constant provides a good description of the data. We also obtain new constraints on the physics of engines that power the large-scale radio emission. The equation that describe the predicted size of each radio source is controlled by one model parameter, $\\beta$, which parameterizes the extraction of energy from the black hole. Joint fits of radio galaxy and supernova samples indicate a best fit value of $\\beta$ that is very close to a special value for which the relationship between the braking magnetic field strength and the properties of the spinning black hole is greatly simplified, and the braking magnetic field strength depends only upon the spin angular momentum per unit mass and the gravitational radius of the black hole. The best fit value of $\\beta$ of 1.5 indicates that the beam power $L_j$ and the initial spin energy of the black hole $E$ are related by $L_j \\propto E^2$, and that the relationship that might naively be expected for an Eddington limited system, $L_j \\propto E$, is quite clearly ruled out for the jets in these systems. ", "introduction": "Recent cosmological studies using CMBR, supernovae, and other types of astronomical sources and phenomena have greatly improved our understanding of the recent expansion and acceleration history of the universe. Whereas a consistent picture has emerged (the ``concordance cosmology''), in which the dynamics of the universe is currently dominated by a mysterious dark energy, its physical nature remains one of the key outstanding problems of physical science today. For a summary of developments in this field see Ratra \\& Vogeley (2007). Aside from the CMBR, most experimental methods to study the expansion history of the universe, and thus its matter-energy contents, require samples of standardisable sources whose distances can be determined in a consistent way, e.g., the supernovae of type Ia. It is clear that both low and high redshift sources play an important role in these studies. Low to moderate redshift sources allow us to define and probe the acceleration of the universe and the properties of the dark energy at the current epoch, whereas higher redshift sources allow us to probe the properties of the dark energy at earlier epochs and possible changes in its properties, which is perhaps the best path towards understanding its physical nature. Confidence in luminosity and coordinate distance determinations, which are the foundation of the studies discussed here, is bolstered when more than one method yields the same results. Like supernovae, powerful radio galaxies are observed out to redshifts greater than one, and their observable properties can be used to determine the coordinate distances to them, or equivalently the luminosity distances or distance moduli (Daly 1994). It is thus interesting to study these sources and to determine whether cosmological results obtained with radio galaxies agree with those obtained with supernovae and other methods. In addition, the radio galaxy and supernova samples may be analyzed jointly to improve the determinations of the radio galaxy model parameters. In this paper, eleven new radio galaxies are combined with a previously studied sample of nineteen sources, to yield a sample of thirty radio galaxies suitable for cosmological studies. This sample is analyzed here in three cosmological models. The first model allows for quintessence and non-relativistic matter in a spatially flat universe; the second model allows for non-relativistic matter, a cosmological constant, and space curvature; and the third model allows for a rolling scalar field in a spatially flat universe (Peebles \\& Ratra 1988). All of these models are based on the equations of General Relativity, and rely upon this as the correct theory of gravity. The primary objectives are to obtain and compare constraints on model and cosmological parameters using different samples of type Ia supernovae, extended radio galaxies, and a combined sample of supernovae and radio galaxies. Similar results obtained with the radio galaxy and supernovae methods, which determine similar quantities over similar redshifts, will bolster our confidence in each method. This is because the methods rely upon measurements of entirely different quantities that are then applied using completely different astrophysical arguments, and thus provide independent measures of coordinate distances and cosmological parameters. Results obtained from the combined sample allow strong constraints to be placed on the radio galaxy model parameter, which provides a direct link to and diagnostic of the physics of the energy extraction that produces the large-scale jets in these systems. Finally, if very similar best fit model parameters are obtained in different cosmological scenarios, tben these parameters can be used to determine the dimensionless coordinate distance to each source. The dimensionless coordinate distances thus obtained are then used for a separate study. In addition, the dimensionless coordinate distances to the radio galaxies can be used to define the distance modulus to each radio galaxy, which can be combined with those of supernovae to increase the sample sizes. The use of powerful extended radio galaxies as a cosmological tool is reviewed in section 2. Results obtained with radio galaxies, supernovae, and combined samples of radio galaxies and supernovae are presented in section 3. Determinations of distance moduli to radio galaxies are presented in section 4. The main results and conclusions of the paper are summarized in section 5. ", "conclusions": "A sample of thiry radio galaxies was used to determine cosmological parameters and the radio galaxy model parameters $\\kappa_{RG}$ and $\\beta$ in three standard cosmological models. Nearly identical values of $\\kappa_{RG}$ and $\\beta$ are obtained in each cosmological model indicating that they are not strongly affected by the context in which they are determined (see Tables 1, 2, and 3), thus they can be used to determine the dimensionless coordinate distance to each source. Three supernovae samples are considered both separately and jointly with the radio galaxy sample and are analyzed in the context of three standard cosmological models (see Tables 1, 2, and 3) to determine cosmological parameters and the model parameter $\\kappa_{SN}$. Nearly identical values of $\\kappa_{SN}$ are obtained for each supernovae sample in the context of each cosmological model, thus they can be used to determine the dimensionless coordinate distance to each source. They can also be used to determine the effective distance modulus to each of the radio galaxies, which can then be combined with those of the supernovae to increase the sample sizes, particularly at high redshift. Constraints on cosmological parameters obtained with radio galaxies are consistent with, though weaker than, those obtained with supernovae alone. There are no inconsistencies between results obtained with radio galaxies and supernovae. For example, in the context of a standard quintessence model, radio galaxies alone indicate that the universe is accelerating today with about 90 \\% confidence (see Figure 6). The consistency between results obtained with radio galaxies alone, supernovae alone, and the combined supernovae and radio galaxy samples suggests that neither method is plagued by unknown systematic errors; both methods seem to be working well. The radio galaxy and supernovae methods are completely independent, based on a completely different physics and observations, and provide independent measures of distances to sources at similar redshift. The facts that the cosmological parameters obtained in specific models are consistent, and that the coordinate distances to sources at similar redshift are consistent suggests that systematic errors are not playing a major role in either method. Since nearly identical values of $\\kappa_{SN}$, $\\kappa_{RG}$, and $\\beta$ are obtained in each cosmological model, they can be used to solve for the dimensionless coordinate distance to each source, $y$. Values of $y$ to each radio galaxy, obtained using the best fit values of $\\kappa_{RG}$ and $\\beta$ indicated by fits to the joint sample of 192 supernovae and 30 radio galaxies, are listed in Table 4. The best fit values of $y$ listed in Table 4 are combined with the best fit value of $\\kappa_{SN}$ obtained for the Davis et al. (2007) sample to obtain the distance modulus to each of the radio galaxies, $\\mu_D$, which are listed in Table 4 and which can be added to those already listed in Davis et al. (2007). Similarly, the values of $y$ to the radio galaxies obtained using the best fit values of $\\kappa_{RG}$ and $\\beta$ indicated by fits to the joint sample of 182 supernovae and 30 radio galaxies were obtained (but are not listed), and were combined with the best fit value of $\\kappa_{SN}$ obtained for the Riess et al. (2007) sample to obtain values of $\\mu_R$ for the radio galaxies that can be combined with those already listed by Riess et al. (2007). Since nearly identical values of $\\kappa_{SN}$, $\\kappa_{RG}$, and $\\beta$ and their uncertainties are obtained in each cosmological model, the average of the values obtained in the different cosmological models and their uncertainties were used. New constraints were obtained on the model parameter $\\alpha$ in the rolling scalar field model of Peebles \\& Ratra (1988). These constraints are rather strong and indicate that $\\alpha$ is close to zero for reasonable values of $\\Omega_m$. Thus, a cosmological constant provides a good description of the data, and there is no indication that the energy density of the dark energy is changing with redshift. New constraints were obtained on the radio galaxy model parameter $\\beta$, suggesting values of $\\beta$ close to 1.5. This is interpreted in the context of a standard magnetic braking model of energy extraction from a rotating black hole (e.g. Blandford 1990). This is a very special value of $\\beta$ for which the braking magnetic field strength depends only upon the spin angular momentum per unit mass and the gravitational radius of the black hole. This suggests that when the magnetic field strength reaches this maximum or limiting value, the relativistic outflow is triggered. The fact that the magnetic field strength does not depend explicitly on the black hole mass for this special value of $\\beta$ may explain why it is that this paricular type of radio source is able to provide a modified standard yardstick for cosmological studies. We have provided further evidence that radio galaxies can be used to determine coordinate distances to sources, and thus cosmological parameters through a standard angular diameter test. Our comparative study shows that values of cosmological parameters obtained with radio galaxies are consistent with those obtained with supernovae. Supernovae alone and radio galaxies alone both indicate that the universe is accelerating at the current epoch when these data are analyzed in specific models such as a a quintessence model in spatially flat universe, a lambda model in universe that allows for non-zero space curvature, and a rolling scalar field model in a spatially flat universe. All of these models rely upon the equations of General Relativity, and the results obtained in these models would not be correct if General Relativity is not the correct theory of gravity. These data are analyzed in a model-independent way that does not rely upon General Relativity by Daly et al. (2008), who show that the supernovae data alone and the radio galaxy data alone indicate that the universe is accelerating at the current epoch independent of whether General Relativity is the correct theory of gravity. When expressed as coordinate distances, both radio galaxy and supernova samples can be combined, and used in a joint cosmological analysis, as it was done, e.g., by Daly \\& Djorgovski (2003, 2004). We use these expanded and combined samples in a separate paper (Daly et al. 2008)." }, "0710/0710.2397_arXiv.txt": { "abstract": "{The initial-final mass relationship (IFMR) for stars is important in many astrophysical fields, such as the evolution of galaxies, the properties of type Ia supernovae (SNe Ia) and the components of dark matter in the Galaxy.} {The purpose of this paper is to obtain the dependence of the IFMR on metallicity.} {Following Paczy\\'{n}ski \\& Zi\\'{o}lkowski (1968) and Han et al. (1994), we assume that the envelope of an asymptotic giant branch (AGB) or a first giant branch (FGB) star is lost when the binding energy of the envelope is equal to zero ($\\Delta W=0$) and the core mass of the AGB star or the FGB star at the point ($\\Delta W=0$) is taken as the final mass. Using this assumption, we calculate the IFMRs for stars of different metallicities.} {We find that the IFMR depends strongly on the metallicity, i.e. $Z=0.0001, 0.0003, 0.001, 0.004, 0.01, 0.02, 0.03, 0.04, 0.05, 0.06, 0.08$ and $0.1$. From $Z=0.04$, the final mass of the stars with a given initial mass increases with increasing or decreasing metallicity. The difference of the final mass due to the metallicity may be up to 0.4 $M_{\\odot}$. A linear fit of the initial-final mass relationship in NGC 2099 (M37) shows a potential evidence of the effect of metallicity on the IFMR. The IFMR for stars of $Z=0.02$ obtained in the paper matches well with those inferred observationally in the Galaxy. For $Z\\geq 0.02$, helium WDs are obtained from the stars of $M_{\\rm i}\\leq 1.0 M_{\\odot}$ and this result is upheld by the discovery of numerous low-mass WDs in NGC 6791 which is a metal-rich old open cluster. Using the IFMR for stars of $Z=0.02$ obtained in the paper, we have reproduced the mass distribution of DA WDs in Sloan DR4 except for some ultra-massive white dwarfs.} {The trend that the mean mass of WDs decreases with effective temperature may originate from the increase of the initial metallicities of stars. We briefly discuss the potential effects of the IFMR on SNe Ia and at the same time, predict that metal-rich low-mass stars may become under-massive white dwarfs.} ", "introduction": "\\label{sect:1} White dwarfs (WDs) are the endpoint of the evolution of stars with initial masses ranging from about 0.1 $M_{\\odot}$ to about 8 $M_{\\odot}$. The vast majority of stars in the Galaxy belong to the mass range and over 97\\% of the stars in the Galaxy will eventually end up as WDs (Fontaine et al. \\cite{FON01}). As a result, WDs may give direct information about star formation during the Galaxy's earliest epochs (Gates et al. \\cite{GAT04}). WDs show their importance in many fields. It has been about 50 yrs since Schmidt (\\cite{SCH59}) recognized the usefulness of WDs as cosmochronometers. WDs may also be the dominating component of dark matter in the Galaxy (Alcock et al. \\cite{ALC99}; Chabrier \\cite{CHA99}). As a matter of fact, WDs are very important for type Ia supernovae (SNe Ia) since it is believed that SNe Ia are from the thermonuclear runaway of carbon-oxygen white dwarfs (CO WD) (see the reviews by Hillebrandt \\& Niemeyer (\\cite{HN00}) and Leibundgut (\\cite{LEI00})). It is well known that the masses of WDs are typically of the order of half that of the Sun, while their radii are similar to that of a planet. However, a detailed knowledge of WDs is still unclear in observation and theory (Fontaine et al. \\cite{FON01}; Moroni \\& Straniero \\cite{MOR02}, \\cite{MOR07}). Among all the uncertainties of WDs, the correlation between initial-final mass relationship (IFMR) and metallicitiy is a very important one. It is well known that low metallicity leads to a larger CO WD for a given initial mass with $Z\\leq 0.02$ (Umeda et al. \\cite{UME99a}). However, it is necessary to check the cases of $Z>0.02$ since $Z\\in[0.06-0.1]$ is possible for some ultra-luminous galaxies (Roberts \\& Hynes \\cite{RH94}; Ruiz-Lapuente et al. \\cite{RUI95}; Terlevich \\& Forbes \\cite{TER02}). The IFMR for stars over a large mass range (e.g. 0.8-8 $M_{\\odot}$) is a powerful input to chemical evolution models of galaxies (including enrichment in the interstellar medium) and therefore can enhance our understanding about star formation efficiencies in these systems (Ferrario et al. \\cite{Fer05}; Kalirai et al. \\cite{KAL07b}). The IFMR is also an important input for modelling the luminosity functions of Galactic disk WDs and the cooling sequences of halo clusters, which may directly yield the age of the Galactic disk and halo components (Kalirai et al. \\cite{KAL07b}). Meanwhile, the IFMR represents the mass loss of a star over its entire evolution and it is possible to get some indications about the origin and evolution of hot gas in elliptical galaxies from the IFMR (Mathews \\cite{MAT90}). Since Weidemann (\\cite{WEI77}) showed the first comparison between observations and theoretical predications of the IFMR, many observations gave constraints on the relationship by studying the properties of white dwarfs in open clusters or field white dwarfs, especially during the last decade (Herwig \\cite{HER95}; Reid \\cite{REI96}; Koester \\& Reimers \\cite{KR96}; Finley \\& Koester \\cite{FK97}; Claver et al. \\cite{CLA01}; Williams et al. \\cite{WIL04}; Ferrario et al. \\cite{Fer05}; Williams et al. \\cite{WIL06}). Some empirical relations have also been suggested based on different observations (Weidemann \\cite{WK83}, \\cite{WEI00}; Ferrario et al. \\cite{Fer05}; Williams et al. \\cite{WIL06}, Dobbie et al. \\cite{DOB06b}). At the same time, numerous sets of stellar evolution models have been calculated to study the relationship in theory and to give the IFMRs for stars of different metallicity (Han et al. \\cite{HAN94}; Girardi et al. \\cite{GIR00}). Although great progress has been made in observation and theory, there still exist many uncertainties, i.e. what is the origin of the intrinsic scatter of WD mass or whether there is any dependence of the IFMR on the metallicity (Kalirai et al. \\cite{KAL05}; Williams \\cite{WIL06}). The correlation between the IFMR and the metallicity has been established in theory for many years (Han et al. \\cite{HAN94}; Girardi et al. \\cite{GIR00}), but the evidence of the dependence of the IFMR on the metallicity was not found until 2005 (Kalirai et al. \\cite{KAL05}). Many theoretical calculations showed that the core mass at the first thermal pulse (TP) in the asymptotic giant branch (AGB) may be taken as the final mass (see the review by Weidemann \\cite{WEI00}). However, these theoretical studies only focused on some special metallicities and it is also difficult for some stellar evolution codes to determine which is the first thermal pulse. In the paper, we use a simple but robust method to systemically study the IFMR over a wide metallicity range, i.e. $0.0001\\leq Z\\leq 0.1$. We also briefly discuss the potential effect of the IFMR on SNe Ia. In section \\ref{sect:2}, we describe our model and physical inputs. We give the results in section \\ref{sect:3} and show discussions and conclusions in section \\ref{sect:4}. \\begin{figure*} \\centering \\includegraphics[width=150mm,height=160mm,angle=270.0]{mimf2.ps} \\caption{Initial-final mass relationship (IFMR) for stars of different metallicities. The points of $M_{\\rm f}< 0.5 M_{\\odot}$ represent helium WDs. The thick solid lines are the mass threshold for the second dredge-up (see the text for details) and the thick dotted ones show the transition between carbon-oxygen white dwarfs and oxygen-neon-magnesium ones in our models. The dashed line in the lower panel shows the IFMR for stars of $Z=0.04$ as a comparison. } \\label{Fig1}% \\end{figure*} ", "conclusions": "\\label{sect:4} \\subsection{the uncertainty of the initial-final mass relationship }\\label{subs:4.1} The major uncertainty of the IFMR is from the assumption that the final mass is equal to the core mass when the binding energy of the envelope $\\Delta W=0$ for an AGB/FGB star. Although the consistency between the mass distribution of WDs in the paper and that from Sloan DR4 upholds the assumption, there is no direct observational evidence to verify it. Another widely accepted choice of the final mass is the core mass of a star at the beginning of the thermal-pulse AGB (TPAGB, the first TP or the end of early AGB) (see the review by Weidemann \\cite{WEI00}). From this choice, taking $Z=0.02$ as an example, the general trend of the IFMR is divided into three segments as follows: (i), the core mass is almost constant from 1 $M_{\\odot}$ to 2.5 $M_{\\odot}$, around 0.52 $M_{\\odot}$, which is smaller than that in this paper by about 0.03-0.1 $M_{\\odot}$. (ii), the core mass strongly increases with the initial mass up to 4.0 $M_{\\odot}$. This trend is similar to that in this paper, but the core mass is smaller than that in this paper by 0.03-0.05 $M_{\\odot}$. (iii), the core mass slowly increases with the initial mass up to 6-7 $M_{\\odot}$. The final mass is smaller than that in this paper by 0.1-0.2 $M_{\\odot}$. According to above discussion , the maximum variation of the final mass derived from different assumptions may be as large as 0.2 $M_{\\odot}$. In this paper, we skipped the detailed evolution of TP and simply treated it as an average evolution, which may lose some information on the structural change of the envelope caused by thermal pulse at the early phase of thermally pulsing AGB. We therefore chose a model with $M_{\\rm i}=3.00 M_{\\odot}$ and $Z=0.02$ to examine the influence of thermal pulses on $\\Delta W$ and the result is shown in Fig. \\ref{Fig8}. From the figure, we see that there is a time that $\\Delta W$ changes its sign but returns back immediately during a full amplitude of TP. The change of the sign of $\\Delta W$ is from the decrease of stellar radius which results from the decrease of total luminosity (see panels 3 and 4). This implies that superwind might be periodically interrupted, but the timescale of the interruption is very short ($\\sim100$yr), and then its influence is small. There are many studies focusing on the TPs previously and we briefly summarize them in the following. Generally, the final mass after several TPs is larger than the core mass at the first TP. Forestini \\& Charbonnel (\\cite{FC97}) calculated a model of $M_{\\rm i}=3 M_{\\odot}$ assuming a Reimers mass-loss law with $\\eta$ increasing from 2.5 to 5.0 in the final AGB stage. They obtained $M_{\\rm f}=0.60M_{\\odot}$ after 19 TPs, but the core mass is 0.54$M_{\\odot}$ at the first TP. Dominguez (\\cite{DOM99}) showed that the thermal pulse starts when the core mass is 0.571$M_{\\odot}$ for a star of $M_{\\rm i}=3 M_{\\odot}$, while the final mass increases to $M_{\\rm f}=0.71 M_{\\odot}$ after 26 TPs if the mass loss prescription of Groenewegen \\& de Jong (\\cite{GRO94}) was adopted (Straniero et al. \\cite{STRA97}). If the fairly strong mass-loss law of Bl$\\ddot{\\rm o}$cker (\\cite{BLOC95}) was adopted, the final mass becomes 0.63$M_{\\odot}$ after 20 TPs from 0.53$M_{\\odot}$ at the first TP (Bl$\\ddot{\\rm o}$cker \\cite{BLOC95}). However, the final mass of 0.68 $M_{\\odot}$ is obtained from the model with $M_{\\rm i}=3.0 M_{\\odot}$ and $Z=0.02$ in this paper. {In Fig. \\ref{Fig8}, we see that $\\Delta W=0$ for a star with $M_{\\rm i}=3.00 M_{\\odot}$ and $Z=0.02$ is achieved before TPAGB phase, which might be very important for galactic chemical evolution since it is widely believed that s-process elements are created in TPAGB phase of a low-mass star (Han et al. \\cite{HAN95}; Busso et al. \\cite{BGW99}). Actually, $\\Delta W=0$ for all the models in this paper is achieved at EAGB phase. These results seem to indicate that the s-process elements could not be created. In fact, as mentioned in subsection \\ref{subs:2.1}, it is more likely that $\\Delta W=0$ is only a lower time limit for superwind, i.e. a superwind starts at or soon after $\\Delta W=0$. Therefore, the stars achieving $\\Delta W=0$ at EAGB phase still have opportunities to enter into TPAGB phase, and contribute to s-process elements.} So, the final mass shown in this paper might be different from a real one. As mentioned above, the core mass in 3 $M_{\\rm \\odot}$ models grows by about 0.1 $M_{\\rm \\odot}$ during TPAGB evolution. For a 1.5 $M_{\\rm \\odot}$ star with $Z=0.02$, the final core mass is about 0.03 $M_{\\rm \\odot}$ larger than that in this paper if the mass loss prescription of Groenewegen \\& de Jong (\\cite{GRO94}) was adopted (Dominguez \\cite{DOM99}). The main result in this paper then still holds. In Fig. \\ref{Fig4}, we see that when $M_{\\rm i}<3.5M_{\\odot}$, the IFMR in this paper is well consistent with the empirical relations. However, for the higher initial mass, the IFMR in this paper is larger than the empirical relations given by Weidemann (\\cite{WEI00}) and Ferrario et al. (\\cite{Fer05}). This discrepancy is mainly from our assumption. As mentioned in subsection \\ref{subs:2.1} and above paragraph, $\\Delta W\\geq0$ is a necessary condition to eject the envelope of an AGB star. Since $\\Delta W=0$ is achieved at the second dredge-up for the star with $M_{\\rm i}>3.5M_{\\odot}$ during EAGB phase and the core is decreasing at this phase, the necessary condition means that final mass may be overestimated. The uncertainty of $M_{\\rm f}$ attributed to the definition of core in this paper may also reduce the discrepancy. Meanwhile, as mentioned in subsection \\ref{subs:3.1}, the final mass might be overestimated if the helium envelope of a star reaches $\\Delta W=0$ again after the ejection of hydrogen-rich envelope. This overestimate is more likely for high initial mass, since the helium envelope is thicker, and then reaches $\\Delta W=0$ more easily (see subsection \\ref{subs:3.1}). However, since the errors from the observations are also very large (see the error bar in Fig. \\ref{Fig4}), the IFMR obtained in this paper is well located in the error range and therefore, is still consistent with observations. \\subsection{low-mass white dwarf}\\label{subs:4.2} Kalirai et al. (\\cite{KAL07a}) suggested that helium white dwarfs may be obtained in metal-rich old clusters, i.e. NGC 6791. It is a reasonable assumption that the same mechanism in NGC 6791 would work for metal-rich field stars. Recently, Kilic, Stanek \\& Pinsonneault (\\cite{KIL07c}) rehandled the data in Valenti \\& Fischer (\\cite{VF05}) and found that there have been metal-rich stars with [Fe/H]$>$ 0 at all times in the local Galactic disk, although the metallicity distribution of disk stars peaks below solar metallicity for stars with ages greater than about 5 Gyrs. Considering that only 5\\% of all WDs in the local disk are single low-mass white dwarfs ($<$ 0.45$M_{\\odot}$) (Liebert et al. \\cite{LIB05}; Kepler et al. \\cite{KEP06}) and the fraction of metal-rich stars with ages greater than 9 Gyrs is 21\\%, Kilic, Stanek \\& Pinsonneault (\\cite{KIL07c}) argued that only the stars of [Fe/H]$>$+0.3 can lose their hydrogen-rich envelope to produce helium WDs on the FGB. From our study, however, all stars with $M_{\\rm i}\\leq1.0 M_{\\odot}$ and $Z\\geq0.02$ will lose their hydrogen-rich envelope and finally become helium white dwarfs. This inconsistency may result from the overestimate of metal-rich stars in the sample of Valenti-Fischer (Reid et al. \\cite{REI07}) adopted by Kilic, Stanek \\& Pinsonneault (\\cite{KIL07c}) As shown in our study, there should be numerous He WDs in metal-rich old clusters, which has indeed been observed in NGC 6791 (Kalirai et al. \\cite{KAL07a}). Here, we emphasize that we do not assume any mass-loss mechanism while Kalirai et al. (\\cite{KAL07a}) assumed a metal-enhanced stellar wind on the red giant branch (RGB). Our study shows that $\\Delta W = 0$ in several FGB stars with high metallicities and small initial masses can be achieved before these stars get to the tip of FGB, indicting that the relative number of the tip RGB stars in metal-rich old clusters is smaller than that in metal-poor clusters. This fact is directly observed in the RGB luminosity functions of two open clusters, e.g. NGC 188 and NGC 6791 (see Fig 8 in Kalirai et al \\cite{KAL07a}). Meanwhile, a similar effect should be seen in the local population of RGB stars. Luck \\& Heiter (see Fig 9 in their paper, \\cite{LH07}) make a comparison of the metallicity histograms between dwarfs and giants within 15 pc of the Sun and found that the dwarf population has a high metallicity tail extending up to [Fe/H]$\\sim$ 0.6, while the giants show a sharp drop in numbers after [Fe/H]=0.2 and no giants with [Fe/H]$>$ 0.45 are observed in the field. The fate that low-mass metal-rich stars will become He WDs before helium is ignited on FGB might give some constrain on planetary nebulae (PNe). It is widely accepted that PNe originate from AGB stars or are related to binary evolution. For the case of binary evolution, one component in the binary system fills its Roche lobe at AGB phase and the mass transfer is dynamically unstable which leads to a common envelope phase. After the ejection of the common envelope, a PN may form. However, since some metal-rich stars may not experience AGB phase, we might speculate that the number of PNe in metal-rich galaxies is relatively lower than that in metal-poor galaxies. Interestingly, Buzzoni et al (\\cite{BUZ06}, see their Fig 11) really observed a relatively low number of PNe per unit galaxy luminosity in more metal-rich elliptical galaxies. Gesicki \\& Zijlstra (\\cite{GZ07}) compared the mass distribution of the central stars of planetary nebulae (CSPN) with those of WDs. These two distribution are very different, i.e. the CSPN mass distribution is sharply peaked at 0.61 $M_{\\odot}$ ranging from 0.55 $M_{\\odot}$ to 0.66 $M_{\\odot}$, while the WD distribution peaks at a slightly lower mass and shows a much broader range of masses. Gesicki \\& Zijlstra (\\cite{GZ07}) suggested that this difference may imply that only some WDs have gone through the PN phase. Our models provide a channel for WDs to avoid the PN phase. Based on our results, WDs from single stars are always larger than 0.4 $M_{\\odot}$, whatever the composition of the WD is. Then, extremely low-mass WDs ($~0.2 M_{\\odot}$) are possible in binary systems or once in binary systems since no stellar population is old enough to produce such extremely low-mass WDs through single star evolution. Only recently, several extremely low-mass ($~0.2 M_{\\odot}$) WDs were discovered in the field (Liebert et al. \\cite{LIB04}; Kawka et al. \\cite{KAW06}; Eisenstein et al. \\cite{EIS06}; Kilic et al. \\cite{KIL07a}), and some of them are companions of pulsars (van Kerkwijk et al. \\cite{VANK05}). Kilic et al. (\\cite{KIL07b}) also found a low-mass WD in a binary system. If an extremely low-mass WD were a single star, it is very likely that it could have a very high space velocity since it may come from a close double-degenerate binary, where its companion has gone through a supernova event that disrupted the binary (Hansen \\cite{HANS03}). LP 400-22 might be a case from this channel (Kawka et al. \\cite{KAW06}). In any case, hard work is need to systematically search for the companions of WDs with mass less than 0.4 $M_{\\odot}$. \\subsection{the potential effect of the initial-final mass relationship on type Ia supernovae}\\label{subs:4.3} As the best cosmological distance indicators, Type Ia supernovae (SNe Ia) have been successfully applied to determine the cosmological parameters ,e.g. $\\Omega_{\\rm M}$ and $\\Omega_{\\rm \\Lambda}$ (Riess et al. \\cite{REI98}; Perlmutter et al. \\cite{PER99}). Phillips relation (Philips \\cite{PHI93}) is used when taking SNe Ia as the distance indicators. It is assumed that Phillips relation is correct at high redshift, although the relation was obtained from a low-redshift sample. This assumption is precarious since the exact nature of SNe Ia is still unclear, especially the progenitor model and explosion mechanism (Hillebrandt \\& Niemeyer \\cite{HN00}; Leibundgut \\cite{LEI00}). If the properties of SNe Ia evolve with redshift, the results for cosmology might be different. Since metallicity decreases with redshift, a good way to study the correlation between the properties of SN Ia and redshift is to study the correlation between the properties of SN Ia and metallicity. It is widely believed that SNe Ia are from thermonuclear runaway of a carbon-oxygen white dwarf (CO WD) in a binary system. The CO WD accretes material from its companion to increase its mass. When its mass reaches its maximum stable mass, it explodes as a thermonuclear runaway and almost half of the WD mass is converted into radioactive nickel-56 (Branch \\cite{BRA04}). The mass of nickel-56 determines the maximum luminosity of SNe Ia. The higher the mass of nickel-56 is, the higher the maximum luminosity is (Arnett \\cite{ARN82}). Some numerical and synthetical results showed that metallcity may affect the final amount of nickel-56, and thus the maximum luminosity of SNe Ia (Timmes et al. \\cite{TIM03}; Travaglio et al. \\cite{TRA05}; Podsiadlowski et al. \\cite{POD06}). There is also much evidence about the correlation between the properties of SNe Ia and metallicity in observations (Branch \\& Bergh \\cite{BB93}; Hamuy et al. \\cite{HAM96}; Wang et al. \\cite{WAN97}; Cappellaro et al. \\cite{CAP97}), e.g. the maximum luminosity of SNE Ia is proportional to the metallicity (Shanks et al. \\cite{SHA02}). Two progenitor models of SNe Ia have competed for about three decades. One is a single degenerate model, which is widely accepted (Whelan \\& Iben \\cite{WI73}). In this model, a CO WD increases its mass by accreting hydrogen- or helium-rich matter from its companion, and explodes when its mass approaches the Chandrasekhar mass limit. The companion may be a main-sequence star (WD+MS) or a red-giant star (WD+RG) (Yungelson et al. \\cite{YUN95}; Li et al. \\cite{LI97}; Hachisu et al. \\cite{HAC99a}, \\cite{HAC99b}; Nomoto et al. \\cite{NOM99}; Langer et al. \\cite{LAN00}). Hachisu \\& Kato (\\cite{HK03a}, \\cite{HK03b}) suggested that supersoft X-ray sources, which belong to WD+MS channel, may be good candidates for the progenitors of SNe Ia. The discovery of the companion of Tycho's supernova also verified the reliability of the model (Ruiz-Lapuente et al. \\cite{RUI04}; Ihara et al. \\cite{IHA07}). Another progenitor model of SNe Ia is a double degenerate model (Iben \\& Tutukov \\cite{IBE84}; Webbink \\cite{WEB84}), in which a system consisting of two CO WDs loses orbital angular momentum by gravitational wave radiation and merges. The merger may explode if the total mass of the system exceeds the Chandrasekhar mass limit (see the reviews by Hillebrandt \\& Niemeyer \\cite{HN00} and Leibundgut \\cite{LEI00}). In both of the progenitor models, the CO WD which finally explodes as a SN Ia should approach or exceed the Chandrasekhar mass limit. Obviously, a higher mass CO WD may fulfill this situation more easily than a low mass one. According to our results, the mass of a CO WD with a given initial mass will be higher in the circumstance with extremely high metallicity or extremely low metallicity than that in the middle-metallicity circumstance. Metallicity is therefore very relevant to SNe Ia. \\subsection{Conclusions}\\label{subs:4.3} We use the method of Han et al. (\\cite{HAN94}) to calculate the IFMR for stars of different metallicities. The conclusions are as follows. \\begin{enumerate} \\item There is an obvious dependence of the IFMRs on the metallicity and the dependence is not monotone. When $Z=0.04$, the final mass of the CO WD for a given initial mass is smallest. For higher or lower $Z$, the mass of CO WD will be higher. The difference of the final mass derived from different metallicties is up to $0.4 M_{\\odot}$. \\item The initial-final mass relationship for stars of $Z=0.02$ is consistent with observations. \\item For $Z\\geq0.02$, a helium white dwarf is formed from a star of $M_{\\rm i}\\leq 1.0M_{\\odot}$. The final masses of helium WDs for a given initial mass slightly decreases with metallicity. \\item Incorporating the IFMR for stars of $Z=0.02$ in the paper into Hurley's single stellar population synthesis code, we reproduce the mass distribution of DA WDs in Sloan DR4 except for some extra-massive WDs. \\item We reconfirm the discovery of Kalirai et al. (\\cite{KAL05}) that the initial-final mass relationship derived from observation in NGC 2099 might be evidence of the dependence of the IFMR on metallicities and that the metallicity of NGC 2099 may be about $0.01$, although the observational error of white dwarfs in NGC 2099 is large. It should be encouraged to program more accurate observations to find a larger sample of white dwarfs in globular clusters. Such programs may help to confirm the dependence of the initial-final mass relationship on the metallicity. \\item We bring up again Willson's suggestion (Willson \\cite{WIL00}) that the effect of metallicity may be the origin of the phenomenon that the mean mass of WDs decrease with effective temperature. \\end{enumerate}" }, "0710/0710.5262_arXiv.txt": { "abstract": "I share a few reminiscences and observations of 40 years of Pulsars. ", "introduction": " ", "conclusions": "" }, "0710/0710.0392_arXiv.txt": { "abstract": "We report 349 radial velocities for 45 metal-poor field red giant and red horizontal branch stars, with time coverage ranging from 1 to 21 years. We have identified one new spectroscopic binary, HD~4306, and one possible such system, HD~184711. We also report 57 radial velocities for 11 of the 91 stars reported on previously by Carney et al.\\ (2003). All but one of the 11 stars had been found to have variable radial velocities. New velocities for the long-period spectroscopic binaries BD$-1$~2582 and HD~108317 have extended the time coverage to 21.7 and 12.5 years, respectively, but in neither case have we yet completed a full orbital period. As was found in the previous study, radial velocity ``jitter\" is present in many of the most luminous stars. Excluding stars showing spectroscopic binary orbital motion, all 7 of the red giants with estimated $M_{\\rm V}$ values more luminous than $-2.0$ display jitter, as well as 3 of the 14 stars with $-2.0 < M_{\\rm V} \\leq\\ -1.4$. We have also measured the line broadening in all the new spectra, using synthetic spectra as templates. Comparison with results from high-resolution and higher signal-to-noise (S/N) spectra employed by other workers shows good agreement down to line broadening levels of 3 \\kms, well below our instrumental resolution of 8.5 \\kms. As the previous work demonstrated, most of the most luminous red giants show significant line broadening, as do many of the red horizontal branch stars, and we discuss briefly possible causes. The line broadening appears related to velocity jitter, in that both appear primarily among the highest luminosity red giants. ", "introduction": "In a previous paper (Carney et al.\\ 2003 hereafter C2003), we discussed results from over two thousand high-resolution (R = 32,000), low signal-to-noise (10 to 50 per resolution element) spectra of 91 field metal-poor red giant branch (RGB) and red horizontal branch (RHB) stars. Radial velocities were obtained by cross-correlating each observed spectrum with a synthetic spectrum that closely matched the adopted temperature, gravity, and metallicity for the star. Sixteen stars were found to be single-lined spectroscopic binaries, and orbital solutions were presented for 14 of them. Excluding those 14 stars, observations of the stars covered spans from 2956 days to 6670 days, roughly from 8 to 18 years, with an average of 13.7 years. The use of synthetic spectra enabled us to measure line broadening as well as radial velocities. We found some anticipated results as well as some surprises. For example, studies of RGB stars in globular clusters (Gunn \\& Griffin 1979; Mayor \\& Mermilliod 1984; Lupton et al.\\ 1987; Pryor et al.\\ 1988; C\\^{o}t\\'{e} et al.\\ 1996, 2002; Brown \\& Wallerstein 1992; Kraft et al.\\ 1997; Mayor et al.\\ 1997) have shown that the most luminous stars, generally with $M_{\\rm V} \\leq\\ -1.4$, show velocity variability that is unlikely to be associated with orbital motion. About half of the stars studied by C2003 with estimated $M_{\\rm V}$ values more luminous than $-1.4$ also showed such velocity variability, which is often referred to as ``jitter\". The term's ambiguity describes the absence of a well-understood cause of the velocity variability. As expected, the spectroscopic binary frequency, for periods less than 6000 days, was very similar to that found for metal-poor dwarfs and subgiants (Latham et al.\\ 2002; Goldberg et al.\\ 2002). There was some evidence for a dearth of short-period binaries among the RGB and RHB stars, which is not too surprising. Most short period systems are expected to undergo mass transfer when the more massive star begins to expand and fills its Roche lobe. The period does shorten initially, until the donor star's mass becomes less than that of the original secondary star, following which the orbital separation widens and the period increases. Among the fourteen spectroscopic binaries in C2003 with orbital solutions, two systems with small ratios of orbital semi-major axis to estimated primary stellar radius have circular orbits, presumably the result of tidal interactions. These two stars also had higher line broadenings than other RGB stars, presumably due to higher rotational velocities, a consequence of tidally-induced locking of the rotational and orbital periods\\footnote{C2003 referred to line broadening as rotational velocities since the values were determined using rotationally-broadened synthetic spectra. Macroturbulence also contributes to line broadening, so throughout this paper we refer to the more general term, line broadening, and refer to $V_{\\rm broad}$ rather than $V_{\\rm rot}$~sin~$i$.}. The most unexpected result involved the dependence of the measured line broadening on the stars' evolutionary state. For most of the RGB stars, the line broadening was generally smaller than our instrumental resolution (about 8.5 \\kms). But at the highest luminosities, the mean line broadening rose to values as high as about 12 \\kms\\ at $M_{\\rm V} \\approx\\ -2.6$, and perhaps as high as 15 \\kms\\ if our interpretation of the apparent periodicity in the velocity jitter of two stars was interpreted correctly as due to the rotational period. These very high levels of line broadening seemed to rule out macroturbulence. We also discounted transfer of core rotational angular momentum to the surface of the RGB stars because we should have seen significant increases in line broadening where internal mixing manifests itself by changes in atmospheric abundances. Having ruled out alternative explanations, we speculated that a ``spin up\" in outer envelope rotational velocities could have been caused by the absorption of giant planets. This would imply that metal-poor stars might have such planets {\\em and} that they exist at larger orbital distances from the host stars than have been found to date in many metal-rich disk stars. We also observed significant line broadening in the RHB stars. To first order, this makes sense. If the much larger stars near the tip of the RGB stars rotate, then so should their descendents. But perhaps their descendents should be stars whose enhanced rotation helps them shed their outer envelopes, which should lead to core helium-burning stars blueward of the instability strip, known as blue horizontal branch (BHB) stars. Soker (1998) and Siess \\& Livio (1999) argued that absorption of a planet by a luminous red giant could explain the high rotation velocities often found among field BHB stars (Kinman et al.\\ 2000; Behr 2003) and cluster BHB stars (Peterson et al.\\ 1983; Peterson 1985a; Peterson et al.\\ 1995; Cohen \\& McCarthy 1997; Behr et al.\\ 2000a,b; Recio-Blanco et al.\\ 2002). More generally, theorists have struggled to explain the presence of both RHB and BHB stars in clusters whose stars share the same metallicity. Some ``second parameter\" must be at work, presumably leading to either larger (RHB) envelope masses or smaller ones (BHB stars). If rotation is the second parameter, then why do RHB stars in the field appear to have rotational velocities comparable to BHB stars, when allowance is made for their different radii? C2003 suggested four follow-up studies. First, expand the sample to ascertain if our original sample was unusual in some manner. This paper reports on the results of a ``hasty\" study of 45 additional metal-poor field stars\\footnote{By ``hasty\", we mean that some stars were observed for less than two years.}. We also sought to obtain line broadening measures for giants in globular clusters, and initial data for four clusters are in hand. Results will be published later. If planets do exist around metal-poor stars, this would suggest that disk instability is a viable mechanism for planet formation, and a high-precision radial velocity survey of roughly two hundred metal-poor field stars was begun (see Sozzetti et al.\\ 2006). Finally, we have to ask if the line broadening is due to rotation or to macroturbulence. This can be determined with very high-resolution, high-S/N spectra that enable Fourier transform studies of line profiles, following Gray (1982; 1984) and Gray \\& Toner (1986; 1987). We have completed acquisition and analysis of such spectra and will report on the results in a future paper (Carney et al.\\ 2007). ", "conclusions": "We have obtained 349 new radial velocities and line broadening measures for 45 metal-poor RGB and RHB stars, as well as 57 such measures for 11 of the stars we studied previously (C2003). A comparison of our derived values for line broadening with results from Behr (2003) and de~Medeiros et al.\\ (2006) shows that our lower-resolution, lower-S/N, and limited wavelength coverage spectra yield excellent results. We believe the good agreement testifies to the power of high-resolution synthetic spectra as templates. We have identified one new spectroscopic binary, HD~4306, and a possible second one, HD~184711, although we note that the latter's radial velocity variability may be due to velocity jitter rather than orbital motion. We draw attention to the observed correlation between variable radial velocity (``jitter\") and line broadening. The significant line broadening seen in the metal-poor field RHB stars is hard to explain: it may be a combination of ``spin up\" when a slowly rotating luminous RGB star settles on to the horizontal branch, or temperature-dependent macroturbulence may be involved. Very high-resolution and very high-S/N spectroscopy should reveal the relative balance of rotation and macroturbulence in RGB and RHB stars." }, "0710/0710.1866_arXiv.txt": { "abstract": "We study singularities which can form in a spherically symmetric gravitational collapse of a general matter field obeying weak energy condition. We show that no energy can reach an outside observer from a null naked singularity. That means they will not be a serious threat to the Cosmic Censorship Conjecture (CCC). For the timelike naked singularities, where only the central shell gets singular, the redshift is always finite and they can in principle, carry energy to a faraway observer. Hence for proving or disproving CCC the study of timelike naked singularities will be more important. Our results are very general and are independent of initial data and the form of the matter. ", "introduction": " ", "conclusions": "" }, "0710/0710.1327_arXiv.txt": { "abstract": "{Despite the enormous progress occurred in the last 10 years, the Gamma-Ray Bursts (GRB) phenomenon is still far to be fully understood. One of the most important open issues that have still to be settled is the afterglow emission above 10 keV, which is almost completely unexplored. This is due to the lack of sensitive enough detectors operating in this energy band. The only detection, by the \\sax/PDS instrument (15-200 keV), of hard X-ray emission from a GRB (the very bright GRB\\,990123), combined with optical and radio observations, seriously challenged the standard scenario in which the dominant mechanism is synchrotron radiation produced in the shock of a ultra-relativistic fireball with the ISM, showing the need of a substantial revision of present models. In this respect, thanks to its unprecedented sensitivity in the 10--80 keV energy band, \\simbolx{}, through follow--up observations of bright GRBs detected and localized by GRB dedicated experiments that will fly in the $>$2010 time frame, will provide an important breakthrough in the GRB field. ", "introduction": "Gamma--Ray Bursts (GRBs) are short and intense flashes of low--energy gamma--rays coming from random directions in the sky at unpredictable times and with a rate of $\\sim$ 300/year as measured by all--sky detectors in low Earth orbit. In the last 10 years, observations allowed huge steps forward in the comprehension of these phenomena, such as their cosmological distance scale, their huge luminosities, their host galaxies, the likely association of \"long\" ($\\sim$2--1000 s) GRBs with the collapse of peculiar massive stars and of \"short\" ($<$$\\sim$2 s) GRBs with the merging of compact objects (NS---NS, NS--BH). See, e.g., \\citet{Meszaros06} for a recent review. However, there are still several open issues, one of the most important being the emission mechanisms in play and their relative contribution to the total radiation. Follow--up observations at longer wavelengths (X--ray, optical, radio) of GRB fields generally lead to the detection of delayed, fading emission (the afterglow). According to the general interpretation, the afterglow emission is described reasonably well, in the framework of the fireball model \\citep{Cavallo78,Meszaros97}, as synchrotron emission from accelerated electrons when a relativistic shell collides with an external medium, the interstellar medium in our case. In this scenario the afterglow spectrum at any given time consists generally of a four segments power law. The spectral and temporal indices are linked together by relationships that depend on the geometry of the fireball expansion \\citep{Sari98} and the properties of the circum--burst environment (density, distribution). The average temporal decaying index,$\\sim$1.3, and spectral photon index, $\\sim$2.2, obtained from observations \\citep{Depasquale06} give an electron spectral index p$\\sim$2.2$-$2.5, which is indeed typical of shock acceleration. \\begin{figure*}[t!] \\centerline{\\includegraphics[clip=true,width=10cm]{amati_f1.eps} } \\caption{\\footnotesize Simulated \\simbolx{} 15--60 keV image of a bright afterglow (like GRB\\,990123) observed for 100 ks starting from 48 hrs since the GRB onset. } \\label{image} \\end{figure*} ", "conclusions": "Despite the enormous observational progress occurred in the last 10 years, the GRB phenomenon is still far to be fully understood. One of the main open issues is the understanding of physical mechanisms at the basis of prompt and afterglow emission; the case of GRB\\,990123 shows that measurements of the nearly unexplored GRB hard ($>$ 15 keV) X-ray afterglow emission can provide very stringent test to emission models. Thanks to its unprecedented sensitivity in the 15--60 keV energy band, \\simbolx{} can provide a significant step forward in this field. Simulations based on observed distribution of X--ray afterglow fluxes and spectral and decay indices show that even a 100 ks TOO observation of a bright GRB starting 2 days after the event can provide sensitive spectral measurements and allow to discriminate different emission components for a significant fraction of events. Moreover, it is likely that significantly lower TOO stat times (12/24 hr) will be possible for a few event/year, thus allowing sensitive hard X--rays measurements also for medium intensity GRB afterglows. The needed GRB detection and few arcmin localizations will be provided by space missions presently planned to be in flight in the $>$2012 time frame and by optical telescopes." }, "0710/0710.3114_arXiv.txt": { "abstract": "The Aarhus code is the result of a long development, starting in 1974, and still ongoing. A novel feature is the integration of the computation of adiabatic oscillations for specified models as part of the code. It offers substantial flexibility in terms of microphysics and has been carefully tested for the computation of solar models. However, considerable development is still required in the treatment of nuclear reactions, diffusion and convective mixing. ", "introduction": "What has become ASTEC started its development in Cambridge around 1974. The initial goal was to provide an improved equilibrium model for investigations of solar stability, following earlier work by \\citet{Christ1974}. However, with the initial evidence for solar oscillations and the prospects for helioseismology \\citep{Christ1976} the goals were soon extended to provide models for comparison with the observed frequencies. Given the expected accuracy of these frequencies, and the need to use them to uncover subtle features of the solar interior, more emphasis was placed on numerical accuracy than was perhaps common at the time. The code drew some inspiration from the Eggleton stellar evolution code \\citep{Egglet1971}, which had been used previously, but the development was fully independent of that code. An early description of the code was given by % \\citet{Christ1982}, with further extensive details provided by \\citet{Christ1978}; many aspects of this still stand and will only be summarized here. With the increasing quality and extent of the helioseismic data the code was further developed, to allow for more realistic physics; a major improvement was the inclusion of diffusion and settling \\citep{Christ1993}. This led to the so-called Model~S of the Sun \\citep{Christ1996} which has found extensive use as reference for helioseismic investigations and which at the time provided reasonably up-to-date representations of the physics of the solar interior. In parallel, extensions have been made to the code to consider the evolution of stars other than the Sun; these include the treatment of convective cores and core overshoot, attempts to model red-giant evolution and the inclusion of helium burning. This development is still very much under way. For use in asteroseismic fitting a version of the code has also been developed in the form of a subroutine with a reasonably simple calling structure which also includes the computation of adiabatic oscillation frequencies as part of the computation. The combined package is available as a single tar file, making the installation relatively straightforward, and the code has been successfully implemented on a variety of platforms. Nevertheless, it is sufficiently complex that a general release is probably not advisable. ", "conclusions": "No stellar evolution code is probably ever finished or fully tested. ASTEC has certainly proved useful in a number of applications, and the results for the Sun, as applied to helioseismology, are perhaps reasonably reliable, at least within the framework of `standard solar modelling'. It is obvious, however, that application to the increasingly accurate and detailed asteroseismic data will require further development. The tests provided through the ESTA collaboration and extended through the HELAS Coordination Action are certainly most valuable in this regard." }, "0710/0710.5732_arXiv.txt": { "abstract": "{} {We performed an integral field spectroscopic study for the HII galaxy IIZw70 to investigate the interplay between its ionized interstellar medium (ISM) and the massive star formation.} {Observations were taken in the optical spectral range from $\\lambda$3700~\\AA~-~6800~\\AA~with the Potsdam Multi-Aperture Spectrophotometer (PMAS) attached to the 3.5 m telescope at Calar Alto Observatory. We created and analyzed maps of spatially distributed emission-lines (at different stages of excitation), continuum emission, and properties of the ionized ISM (ionization structure indicators, physical-chemical conditions, dust extinction, kinematics). We investigated the relation of these properties to the spatial distribution and evolutionary stage of the massive stars.} {For the first time we have detected Wolf-Rayet (WR) stars in this galaxy. The peak of the ionized gas emission coincides with both the location of the maximum of the stellar continuum emission and the WR bump. The region of the galaxy with lower dust extinction corresponds to the region that shows the lowest values of velocity dispersion and radial velocity. The overall picture suggests that the ISM of this region is being disrupted via photoionization and stellar winds, leading to a spatial decoupling between gas+stars and dust clouds. The bulk of dust appears to be located at the boundaries of the region occupied by the probable ionizing cluster. We also found that this region is associated to the nebular emission in HeII$\\lambda$4686 and to the intensity maximum of most emission lines. This indicates that the hard ionizing radiation responsible for the HeII$\\lambda$4686 nebular emission can be related to the youngest stars. Within $\\sim$ 0.4 $\\times$ 0.3 kpc$^{2}$ in the central burst, we derived oxygen abundances using direct determinations of $T_{\\rm e}$[OIII]. We found abundances in the range 12+log(O/H)= 7.65 - 8.05, yielding an error-weighted mean of 12+log(O/H)=7.86 $\\pm$ 0.05 that has been taken as the representative oxygen abundance for IIZw70.} {} ", "introduction": "\\label{intro} The HII galaxies are gas-rich, dwarf systems that have experienced intense recent or ongoing violent star formation (SF). These objects were identified for the first time by Haro (1956) and Zwicky (1964), and are characterized by the presence of giant HII regions that dominate their observable properties at optical wavelengths. In particular, their integrated spectra are very similar to those of giant extragalactic HII regions in spiral galaxies (Sargent \\& Searle 1970). The analysis of their emission line spectrum indicates that HII galaxies are metal poor objects (1/40 Z$_{\\odot}$ $\\lesssim$ Z $\\lesssim$ 1/3 Z$_{\\odot}$; e.g. Kunth \\& Sargent 1983). HII galaxies are ideal laboratories for probing the interplay between massive SF and the ISM in low metallicity environments. The massive SF process in dwarf galaxies has effects on the properties of the surrounding ISM. A large number of massive stars (between 10$^{4}$ - 10$^{6}$ solar masses of gas are transformed into stars in HII galaxies) are formed almost simultaneously within relatively small volumes (e.g. Melnick 1992). Millions of years after the onset of the burst, the most massive stars begin to explode as supernovae, creating violent, short-lived injections of kinetic energy and metallic elements into the ISM. The disruption of the ISM may significantly affect the spatial distribution of gas and dust particles in the regions close to the massive star cluster, and also determine the way in which new metals ejected by massive stars are mixed with the original gas from which the stars formed (Kunth \\& \\\"Ostlin 2000). High spatial resolution imaging has revealed that in many HII galaxies the ionized material presents a complex structure: star clusters in the main body (with a non-uniform distribution of SF knots, ensembles of star clusters, or individual super stellar clusters) of the galaxy and external gas components outside the young stellar clusters (e.g. Martin \\& Kennicutt 1995, V\\'{\\i}lchez \\& Iglesias-P\\'aramo 1998, Papaderos et al. 2002). Thus, a two-dimensional analysis of the ionized material in HII galaxies yields spatially resolved information on properties of the ionized gas. The study of the distribution of these properties is an important issue for our understanding of the interplay between the massive stellar population and the ISM. However, up to now only a few two-dimensional spectroscopic studies of HII galaxies have been performed. Recent work by Izotov et al. (2006) presents two-dimensional spectroscopy of the extremely metal-deficient HII galaxy SBS 0335-052E. They found a small gradient of the electron temperature $T_e$ and oxygen abundance, and the presence of an ionized gas outflow in the perpendicular direction to the galaxy disk. Cair\\'os et al. (2002), using two-dimensional spectroscopy, obtained the ionized gas velocity field in the central part of Mrk 370. In this paper, we present a two-dimensional spectroscopic study for the HII galaxy IIZw70. This galaxy is number 70 in Zwicky's second list of ``Compact Galaxies and Compact Parts of Galaxies, Eruptive and Post-eruptive Galaxies''(Zwicky \\& Zwicky 1971). IIZw70 has been classified as an HII galaxy (Sargent 1970). The basic data of IIZw70 are shown in table~\\ref{tab1}. This object is interacting with its nearby companion IIZw71 at a projected distance of 23 kpc (assuming a distance to the IIZw70-71 system of 18.1 Mpc; see table 1). The ongoing interaction is indicated by interferometric \\ion{H}{I} studies (Balkowski et al. 1978, Cox et al. 2001) which revealed a gaseous streamer connecting IIZw70 with the polar ring galaxy candidate IIZw71 (Whitmore et al. 1990, Reshetnikov \\& Combes 1994). Cair\\'os et al. (2001) show that IIZw70 presents very elongated outer isophotes and a very blue (U-B = -0.89) nuclear starburst in what seems to be an edge-on disk. They also find that the starburst activity is concentrated in the optical center of the galaxy and is surrounded by faint gaseous emission with quite a distorted morphology. \\begin{figure} \\center{ \\includegraphics[width=9cm,clip]{7987fig1.ps}} \\caption{Image of IIZW70 in the {\\it r (left)} and {\\it u (right)} bands from Sloan Digital Sky Survey. The blue box mosaic represents the field used in this work ({\\it left}), which is located approximately in the central parts of the galaxy; the green box marks the likely ionizing clusters location, as indicated with a box on the {\\it u} image ({\\it right}). North is up, east is left. The spatial scale is shown in each panel, where 1 kpc $\\sim$ 11\\hbox{$^{\\prime\\prime }$} assuming a distance to the galaxy of 18.1 Mpc (see table 1).} \\label{SDSS} \\end{figure} In figure~\\ref{SDSS} we show the images of IIZw70 in the {\\it r} (left image) and {\\it u} (right image) bands extracted from SDSS\\footnote{The Sloan Digital Sky Survey website is http://www.sdss.org}. Over the {\\it r} filter image, we marked the field of view used in this work (blue box mosaic). In this work, we investigate the relation between massive SF process and the ionized ISM in IIZw70 using integral field spectroscopy (IFS). In order to do this we created and analyzed maps of spatially distributed emission-lines, continuum emission and physical-chemical properties of the ionized gas (electron temperature and density, gaseous metal abundances, dust extinction, excitation). We present a discussion on the significance of the chemical abundance variation found here. We also present a brief study of the kinematic properties of the ionized gas. The relationship between the physical-chemical and kinematic properties spatial distribution is discussed. \\begin{table}% \\caption{Basic data of IIZw70 \\label{tab1}} \\renewcommand{\\footnoterule}{} \\begin{minipage}{\\textwidth} \\begin{tabular}{lc} \\hline Parameter &Value \\\\ \\hline Name &IIZw70 \\\\ Other designations &UGC 9560, Mrk 829 \\\\ R.A. (J2000.0) &14h 50m 56.5s \\\\ DEC. (J2000.0) &+35d 34' 18'' \\\\ redshift &0.004 \\\\ M$_{B}$\\footnote{Absolute magnitude in B from Cair\\'os (2001)}(mag) &-16.3 \\\\ m$_{B}$\\footnote{Apparent magnitude in B from Deeg et al. (1997)}(mag) &15.1 \\\\ D\\footnote{Distance to the galaxy from Cox et al. (2001)}(Mpc) &18.1 \\\\ M$_{HI}$\\footnote{Neutral hydrogen mass from Thuan \\& Martin (1981)}(\\msun) &0.34 x 10$^{9}$ \\\\ Z\\footnote{Metallicity from Kobulnicky \\& Skillman (1996)}/Z$_{\\odot}$ & 1/5 \\\\ \\hline \\end{tabular} \\end{minipage} \\end{table} In the following section we describe the observations and data reduction. In Sect.3 the results are presented. We discuss our results in Sect.4. ", "conclusions": "\\subsection{Ionized gas and massive stars} We analyzed integral field spectroscopic observations of the HII galaxy IIZw70 in order to investigate the interplay between the massive stars and the ionized gas. This galaxy is interacting with its nearby companion IIZw71. The contiguous streamer of gas between IIZw70 and IIZw71, as well as the SF activity seen in both, indicate an ongoing interaction and possible gas interchange between these two galaxies (Cox et al. 2001). We studied the spatial distribution of properties of the ionized gas (electron temperature and density, chemical abundances, dust extinction, excitation, kinematics) from different emission-line maps. These maps reveal a central starburst (Cair\\'os et al. 2001, 2001) surrounded by an extended lower excitation, low surface brightness ionized component. From our integrated flux in H$\\alpha$, corrected by extinction, we derived a star formation rate of SFR(H$\\alpha$) = 0.3 M$_{\\odot}$/yr for IIZw70 following Kennicutt (1998); this value is in agreement with the value we derived using the flux reported in Gil de Paz et al. (2003) from CCD H$\\alpha$ imaging. Kewley et al. (2002), using long-slit spectroscopic mapping, derived a SFR(H$\\alpha$) = 0.19 M$_{\\odot}$/yr\\footnote{We calculated this value converting the SFR(H$\\alpha$) given in Kewley et al. (2002) from the distance they adopted to the distance adopted in this paper (see table 1).}. Different integration areas and C(H$\\beta$) values could be possible reasons for the discrepancy between our SFR(H$\\alpha$) measurement and the value found by Kewley et al. (2002). For the first time we detected the presence of WR stars in IIZw70. For the blue bump, our measurements of the equivalent width, EW(WRbump) = -2.2 $\\pm$ 0.6 \\AA, and of the ratio 100 x I(WRbump)/I(H$\\beta$) = 4.6 $\\pm$ 1.1 (section 3.3) are in good agreement with the predictions of the evolutionary synthesis models Starburst99 (assuming a Salpeter IMF, M$_{upper}$ = 120 M$_{\\odot}$ and Z=0.004 with high mass loss; Leitherer et al. 1999). The models predict EW(WRbump) values around -2 \\AA~ at $\\sim$ 3.2 Myr after an instantaneous burst. Corresponding predictions for 100 x I(WRbump)/I(H$\\beta$) give values around 3.6 Myr. This fact indicates that the central burst in IIZw70 is very young. Assuming an age for the burst of 3.2 Myr (as deduced from our measurements), we obtained a ratio of WR/O = 0.056 from Starburst99. The presence of a bump at $\\sim$ 5800 \\AA~ could be an indicator of the existence of WC stars. A broad feature at this wavelength has been marginally detected in the same spectra, and though the measured EW is consistent with the same model predictions, its statistical significance is poor due to the low S/N. Overall, a total number of $\\sim$ 17 WN+WC stars are predicted by the models from our findings. The nebular emission of HeII$\\lambda$4686 detected in this work is shown to be spatially extended, though it is not coincident with the location of the WR bumps by a few arcsec\\footnote{We should bear in mind, however, that a small contribution of nebular HeII$\\lambda$4686 could have remained unnoticed in the broad feature}. This nebular HeII emission is seen to the southwest of the box encompassing the probable ionizing clusters (see figure 4), suggesting the existence of very hard ionizing radiation which, in principle, could not be necessarily related to the WR stars located towards the northeast corner. In the overall picture, IIZw70 is dominated by the presence of a young central starburst with a significant SFR. The peak of the H$\\alpha$ emission is located close to the probable youngest stellar clusters (see figure~\\ref{cumulo.WR.HeII}). These young starburst could be producing strong winds which have been able to develop an ionized gas outflow from which we see the blueshifted component as a feature in the radial velocity map of the galaxy (see figure 10). This outflow is thought to have been responsible for the minimum in optical extinction (slightly displaced from H$\\alpha$ emission peak by $\\sim$2$^{\\prime\\prime }$ to the west) apparent in the H$\\alpha$/H$\\beta$ map, where dust could have been swept up by the wind. The characteristic size ($\\sim$ 150 pc) and velocity ($\\sim$ 60 km/s) of the outflow zone are consistent with previous findings in the literature (see e.g. Roy et al. 1991) for the giant HII complex NGC~2363). We can also see that the area with lower dispersion velocity partially overlaps with the zone of the gas outflow, between the probable ionizing clusters location and the zone of minimum extinction. From the kinematical point of view, this picture appears to be consistent with the velocity dispersion map. The narrowest line profiles are dominated by the bulk of the emission from the central HII regions, as expected (Telles et al. 2001), whereas the rest of the ionized gas, with a lower surface brightness, seems to be suffering the effects of the winds and its kinematics should be associated to the structures generated by the likely interaction of the winds with the ambient medium. This situation is reminiscent of what has been found for other extragalactic HII regions and HII galaxies (e.g. Mu\\~noz-Tu\\~n\\'on et al. 1996, Telles et al. 2001) which host strong starbursts. \\subsection{Chemical enrichment and abundance variations} In this section, based on our chemical abundance measurements, we present an analysis on the chemical enrichment of the ionized gas and chemical abundance variations in IIZw70. The oxygen abundance derived here for the integrated spectrum of IIZw70 is 12 + log(O/H) = 7.83 $\\pm$ 0.04. Previous works reported direct values of oxygen abundance for IIZw70 using long-slit spectra positioned in different places across the brightest regions of IIZw70. Lequeux et al. (1979) quoted a value of 8.07; Kobulnicky \\& Skillmann (1996) derived an oxygen abundance of 8.07 $\\pm$ 0.08 whereas Shi et al. (2005) found 12+log (O/H) = 7.69. In principle, the differences between our O/H measurement and the values found in the literature could be explained by different integration areas (see figure 9, bottom panel; Kewley et al. 2005) and atomic parameters used. Thus far, in most works on chemical abundances in dwarf galaxies there is no clear evidence of abundance variations. Though selected examples of nitrogen enhancement in dwarf galaxies have been reported in the literature, notably the case of NGC~5253 (e.g. L\\'opez-S\\'anchez et al. 2006 and references therein), this is not the case of O/H. Many works in the literature have searched for a possible oxygen abundance variation within dwarf galaxies. Kobulnicky \\& Skillman (1996) presented optical spectroscopic chemical measurements of the ISM in the irregular galaxy NGC 1569 that reveal no substantial localized chemical self-enrichment (i.e., ``local pollution''; see e.g. Kunth \\& Sargent 1986; Pagel, Terlevich \\& Melnick 1986; Pilyugin 1992). V\\'{\\i}lchez \\& Iglesias-P\\'aramo (1998), using bi-dimensional, long-slit spectroscopy, showed that IZw18 presents a substantially homogeneous chemical composition over the whole galaxy. From long-slit spectra of 16 HII regions in the dwarf irregular NGC 1705, Lee \\& Skillman (2004) reported that there is no significant spatial variation of oxygen abundances for this galaxy. Izotov et al. (2006), from two-dimensional spectroscopy of the HII galaxy SBS0335-052E, found a variation of 0.4 dex in the oxygen abundance. They concluded that these variations may not be statistically significant due to unaccounted error estimates. Lee et al. (2006) obtained a slightly significant (3.2$\\sigma$) oxygen abundance gradient of -0.16 $\\pm$ 0.05 dex kpc$^{-1}$ within the Local Group dwarf irregular galaxy NGC~6822. However they claim further deep high-quality spectra of nebulae and stars are needed to distinguish clearly between either a zero or a nonzero slope. In this work we calculated the O/H abundance in 16 positions across IIZw70 directly from the measurement of [OIII]$\\lambda$4363. From these values we found a maximum variation in the derived oxygen abundances of $\\sim$ 0.4 dex, similar to the range reported by Izotov et al. (2006) for SBS0335-052E. In order to study whether this variance is statistically significant for an effective abundance variation we must include all other unaccounted uncertainties (variable seeing, flux calibration, etc), which surely contribute to the total oxygen abundance error but are difficult to estimate. In any case, our observations indicate that, within $\\sim$ $\\pm$ 0.2 dex, the ionized gas in the central burst of IIZw70 appears to be chemically homogeneous in O/H over spatial scales of hundreds of parsecs. In the case of the neon-to-oxygen ratio, Ne/O, the range of values as derived from our point by point abundance analysis, does not show inhomogeneity within the errors; as is also the case of the nitrogen-to-oxygen ratio, N/O (see section 3.4). All these findings give us the upper limits to any chemical inhomogeneity in the ionized gas in the brightest part of IIZw70. In accordance with these results we can say that we are probably not detecting a ``local pollution'' case in IIZw70, consistent with most results obtained by other works cited previously. A favourable explanation to this apparent degree of homogeneity in dwarf galaxies is given by the scenario presented in Tenorio-Tagle (1996). According to this scenario, in those places in which one expects that newly produced metals could be returned to the ISM, the ``local pollution'' is not detected in the warm phase of the ISM. Rather the newly synthesized metals are returned into the hot phase of the ISM, and there is a significant time delay before these newly produced metals can be detected in the warm phase of the ISM. Variations of the oxygen abundance in the ISM of HII galaxies could give constraints to the physical mechanisms involved in the recycling processes (e.g. Recchi et al. 2001). Analysing the predictions of chemical evolution models, we verified that continuous SF models (Moll\\'a \\& D\\'\\i az 2005) predict an enhancement in O/H of up to 0.4 dex, to be produced in spatial scales of $\\sim$ 1 kpc along their disk, over timescales of 8 Gyr. The important point here is that these models predict O/H enhancements that could be achieved keeping the N/O ratio approximately constant throughout the process, and at an absolute value consistent with our observations (see section 3.4) to within 0.1 dex. Further observations with IFUs in 10 m class telescopes and additional measurements of chemical abundances in IIZw70, covering a larger area, should be required in order to draw a firmer conclusion on the statistical contribution of local variations to the chemical homogeneity. Given the interest in the problem of chemical homogeneity of dwarfs galaxies and its role in their chemical evolution (e.g. Legrand et al. 2001, Recchi et al. 2001) a clearer view of the observational situation is probably warranted." }, "0710/0710.0784_arXiv.txt": { "abstract": "Spectroscopic modeling of Type II supernovae (SNe) generally assumes steady-state. Following the recent suggestion of Utrobin \\& Chugai, but using the 1D non-LTE line-blanketed model atmosphere code CMFGEN, we investigate the effects of including time-dependent terms, which are generally neglected, that appear in the statistical and radiative equilibrium equations. We base our discussion on the ejecta properties and the spectroscopic signatures obtained from time-dependent simulations, investigating different ejecta configurations (slow, standard, and fast), and covering their evolution from one day to six weeks after shock breakout. Compared to equivalent steady-state models, our time-dependent models produce SN ejecta that are systematically over-ionized, affecting helium at one week after explosion, but ultimately affecting all ions after a few weeks. While the continuum remains essentially unchanged, time-dependence effects on observed spectral lines are large. At the recombination epoch, H{\\sc i} lines and Na{\\sc i}\\,D are considerably stronger and broader than in equivalent steady-state models, while Ca{\\sc ii}\\,8500\\AA\\ is weakened. If time dependence is allowed for, the He{\\sc i} lines at 5875\\AA\\ and 10830\\AA\\ appear $\\sim$3 times stronger at one week, and He{\\sc i}\\,10830\\AA\\ persists as a blue-shifted absorption feature even at 6 weeks after explosion. Time dependence operates through the energy gain from changes in ionization and excitation, and, perhaps more universally across SN types, from the competition between recombination and expansion, which in-turn, can be affected by optical-depth effects. Our time-dependent models compare well with observations of the low-luminosity low-velocity SN\\,1999br and the more standard SN\\,1999em, reproducing the H$\\alpha$ line strength at the recombination epoch, and without the need for setting unphysical requirements on the magnitude of nickel mixing. ", "introduction": "Because of radiative cooling and the fast expansion of the exploding mantle of the progenitor massive star, photospheric-phase Type II SN spectra evolve rapidly. At shock breakout the spectral energy distribution (SED) peaks in the far-UV and X-rays. Subsequently the SED centers in the UV, then in the optical, and later in the infrared until the object fades from view. From a radiative transfer perspective, the cooling induces a recombination to lower ionization stages that impose a strong blanketing effect on the energy distribution. A few days after explosion, the ejecta radiate a UV-dominated SED, with He{\\sc ii}\\,4686\\AA\\ (only observed at a few days after shock breakout; Dessart et al. 2007), Balmer/Paschen lines of Hydrogen, He{\\sc i}\\,5875\\AA, He{\\sc i}\\,10830\\AA, some multiplets of O{\\sc ii} and N{\\sc ii} multiplets at \\,4600\\AA\\ (see Dessart \\& Hillier 2006a; Baron et al. 2007), N{\\sc ii}\\,5400\\AA, and a few isolated resonance lines such as Mg{\\sc ii}\\,2802\\AA\\ and Al{\\sc iii}\\,1859\\AA. As the ejecta continue to cool, hydrogen eventually recombines, with a contemporaneous enhancement in line-blanketing due to the switch from Fe{\\sc iii} to Fe{\\sc ii}. In Dessart \\& Hillier (2006) and Dessart et al. (2007), we successfully modeled these epochs for SNe 1999em, 2005cs, and 2006bp, using steady-state non-LTE CMFGEN models and a power-law density distribution with an exponent of ten (the immediate post-breakout phase requires higher values for this exponent). These spectroscopic analyses deliberately focused on the first $\\sim$45 days after shock breakout, since, beyond that time, we encountered severe difficulties in reproducing the hydrogen lines. Specifically, at late times, synthetic line profiles were sizably narrower than observed, suggesting that line formation regions predicted by CMFGEN were confined to velocities/radii that were too small. Despite a continued success with reproducing the overall continuum energy distribution, CMFGEN failed to reproduce important line profiles, whenever hydrogen started recombining at and above the photosphere. The H$\\alpha$ problem has not been clearly emphasized in the literature, where one can see a great disparity in model atmosphere assumptions and agreement between theoretical predictions and observations, but no direct link to physical/numerical issues. In the eighties and early nineties, the recognized importance of non-LTE effects confronted the strong limitations of computer technology, so that only a few species were treated in non-LTE (i.e., usually hydrogen and helium), while the metals responsible for line blanketing were treated in LTE. \\citet{Eastman_Kirshner_1989} followed this approach to model the first ten days of SN 1987A, and did not encounter obvious difficulties with hydrogen lines. \\citet{Schmutz_etal_1990}, using an approximate non-LTE technique, had, on the contrary, great difficulty reproducing any of the Balmer lines, suggesting clumping as the culprit. \\citet{Hoeflich_1988} reproduced the SN 1987A spectral evolution and the hydrogen lines, over many months, using a large ``turbulent'' velocity. In the more sophisticated non-LTE CMFGEN \\citep{HM_98,DH_05a} models presented in \\citet{DH_06a}, we found that decreasing the turbulent velocity weakened line-blanketing effects and increased, although only modestly, the strength of hydrogen lines. However, beyond 40 days after explosion, this tuning had no longer any important influence on the hydrogen lines. The non-LTE model atmosphere code PHOENIX predicts strong Balmer lines at the hydrogen recombination epoch \\citep{Mitchell_etal_2001,Baron_etal_2003}, but this may stem from their adoption of non-thermal ionization/excitation due to $^{56}$Ni at the photosphere, sometimes just a few days after explosion and in mass shells moving at $\\ge$10000\\,\\kms in SN 1987A \\citep{Mitchell_etal_2001} or $\\sim$9000\\,\\kms in SN 1993W \\citep{Baron_etal_2003}. Hydrodynamical simulations of core-collapse SNe predict that $^{56}$Ni has velocities of at most $\\sim$4000\\,\\kms, the nickel fingers being strongly decelerated at the H/He interface \\citep{Fryxell_etal_1991,Kifonidis_etal_2000, Kifonidis_etal_2003}. The magnitude of this disagreement extends far beyond the uncertainties of explosion models and suggests a genuine incompatibility. Interestingly, H{\\sc i} lines remain strong for months in all Type II SNe, in objects as diverse as the ``peculiar'' SN 1987A, the ``plateau'' SN 1999em, and the ``low-luminosity'' SN 1999br. CMFGEN models computed for these objects were unable to reproduce Balmer lines after $\\sim$4 days in SN 1987A, $\\sim$40 days in SN 1999em, and $\\sim$20 days in SN 1999br, all coincident with hydrogen recombination in the ejecta. These three SNe have very different inferred ejecta properties and observed light curves. SN 1999br even synthesized an order of magnitude less $^{56}$Ni than average \\citep{Pastorello_etal_2004} for the Type II class. The only common property between these objects, which is connected to the H$\\alpha$ problem, is the recombination of the ejecta to a lower ionization state at the corresponding epoch. Recently, \\citet[UC05]{UC_05} proposed that the effect of time-dependence, and the energy associated with changes in ionization/excitation, lead to a strong H$\\alpha$ line profile in SN 1987A during the recombination epoch. They also found that barium lines were affected, and that, with their more consistent approach, Ba{\\sc ii}\\,6142\\AA\\ could be fitted using the LMC metallicity value. The steady-state models of \\citet{Mazzali_etal_1992} supported instead an abundance enhancement of five. Time dependence has been invoked in the past by \\citet{Fransson_Kozma_1993} to explain the late-time light curve of SN 1987A, and the theoretical study of \\citet{Pinto_Eastman_2000a, Pinto_Eastman_2000b} showed that time dependence {\\it in the radiation field} had a critical impact on the radiative transfer in Type Ias. \\citet{Pinto_Eastman_2000a, Pinto_Eastman_2000b}, however, treat the material in LTE, i.e., do not solve the rate equations, focusing instead on the time-dependent {\\it diffusion} of photons through an optically thick Type Ia SN ejecta. Similarly, \\citet{Kasen_etal_2006} neglect explicit time dependence in the rate equations, computing the ionization and excitation state of the medium in LTE. Thus, although there is at present growing interest in accounting for time dependence in the radiation field, the often-used expedient of LTE, to maintain low CPU costs, has forced the neglect of both non-LTE and time-dependence in the level populations. One exception to these time-dependent LTE approaches is the work of \\citet{Hoeflich_2003}, who treats time-dependence in the rate equations but, to our knowledge, has not discussed the associated effects on Type II SN spectra. In the present study, we investigate thoroughly the effects of time-dependence in the rate equations and their impact on inferred ejecta properties. Note that time dependence in the radiation field is accounted for by adjusting the base luminosity so that the emergent synthetic flux matches the observed flux, using the bolometric-light evolution of SN 1999em as a guide \\citep{DH_06a}. Here, we report the salient features of several {\\it time-dependent} non-LTE CMFGEN simulations, covering the early evolution for a range of Type II SN ejecta. We confirm the results of UC05 that time-dependence induces an over-ionization of the recombining ejecta and that it solves the H$\\alpha$ problem. However unlike UC05, we self-consistently solve the radiation transfer equation --- the coupling between the level populations and the radiation field is calculated and fully allowed for. Our more comprehensive study of the full spectrum further predicts that all lines, not just those of hydrogen, are significantly affected. The ionization of the ejecta and its evolution are so strongly modified that we predict even He{\\sc i}\\,10830\\AA\\ many weeks after explosion. In the next section, we present out treatment of time-dependence in CMFGEN, which focuses here on the terms appearing in the statistical and radiative equilibrium equations (an appendix also provides further details). In \\S2.5, we present the various model calculations performed to illustrate our discussion. In \\S3, we present our results, discussing the effects of time dependence on the ejecta properties as well as the associated spectroscopic signatures. We then present in \\S4 a comparison with a few representative observations, focusing on the well observed Type II Plateau SN 1999em and the low luminosity Type II SN 1999br. In \\S5, we discuss the implications of such time-dependent effects on our understanding and on our modeling of photospheric-phase Type II SN spectra, before giving our conclusions in \\S6. Note that the salient features and key results presented here are also available in a concise, ``letter'', format in \\citet{DH_06b}, to which we refer the hurried reader. ", "conclusions": "We have presented an extension to the non-LTE model atmosphere code CMFGEN to treat time-dependent terms in the statistical and radiative equilibrium equations. We use implicit first order differencing, and the resulting equations are solved by a partial linearization technique in a manner virtually identical to that used to solve the steady state equations \\citep{HM_98}. We confirm the findings of Utrobin \\& Chugai (2005), which were related to H$\\alpha$ and Ba{\\sc ii}\\,6142\\AA\\ at a few days after explosion in the spectrum of SN 1987A, that the time dependent terms are important for the analysis of Type II SN spectra. We find that the inclusion of time-dependence terms in the statistical equilibrium equations produces ejecta, in Type II SN, that are systematically over-ionized relative to steady-state models. Spectroscopically, the associated changes in level populations and optical depths alter the line profiles of all species throughout the spectral range. Qualitatively, lines appear in general stronger and broader, as is seen for the Balmer, the Paschen, and the Brackett series of hydrogen or for Na{\\sc i}\\,D. Exceptions (e.g. Ca{\\sc ii}\\,8500\\AA) arise when the altered ionization inhibits the necessary recombination of a given ion (e.g., Ca$^{++}$). Because of optical depth effects, the inclusion of time-dependent terms in the statistical equilibrium equations can even influence the ionization state of the gas in regions where the classic recombination times-scale is much shorter than the flow time-scale. Surprisingly, we predict the presence of He{\\sc i}\\,10830\\AA\\ 6 weeks after the explosion, naturally caused by the frozen, and full, ionization of the outer ejecta. The resulting detached blueshifted absorption and flat-topped emission profile is supported by observations, although the magnitude of the absorption component may require the additional contribution from the cold dense shell that can form at the interface between the SN ejecta and the pre-SN wind \\citep{Chugai_etal_2007}. The effects of time dependence observed stem from the energy gain that follows changes in ionization and excitation, and from the similarity in recombination and expansion time-scales (crucial at large distances above the photosphere). The recombination time-scale can be lengthened by optical-depth effects, and in some cases this may allow time-dependence effects to be important even at the photosphere. The present study indicates that neglecting time dependence may compromise the results of quantitative analyses of Type II SN spectra, affecting the inferred ejecta ionization (and its origin) or chemical abundances. In particular, time dependence offers a natural means to sustain strong and broad lines at late times, as observed in {\\it all} Type II SN spectra, without invoking radioactive contributions. Such non-thermal excitation/ionization is expected to be relevant at the photosphere after a few months, but not as soon as a few weeks, and is subject to strong variations accross the SN class (SNe 1999br or 2005cs, which boast a strong H$\\alpha$ at the recombination epoch, have symptomatically low $^{56}$Ni yields compared to the standard Type II-Plateau SN 1999em). The time-dependence effects we observe just one week after explosion in the outer, fully ionized, SN ejecta also support the idea that the energy gain drawn out of recombining ions is not a fundamental driver. Time-dependence effects can lead to over-ionization for He, C, N, O, and metals, under fully-ionized hydrogen conditions. We thus surmise that time dependence should operate, with a magnitude to be determined, in SN ejecta of all types, primarily because they all combine the properties of fast expansion and low density. We have investigated the potential impact of time dependence on the correction factors used in the Expanding Photosphere Method, or, essentially, whether the influence on the electron density generates a systematic shift in the magnitude of flux dilution. We find no sizable and systematic deviation from the correction factors obtained, at the same color temperature, with steady-state CMFGEN models. This supports our use of steady-state CMFGEN models for distance determinations based on early-time Type II SN observations (Dessart \\& Hillier 2006a; Dessart et al. 2007). However, the good agreement at late times between time-dependent CMFGEN models and late time photospheric-phase observations motivates the extension of the time baseline to include the recombination epoch, thus allowing for multi-epoch observations that cover the entire Plateau phase, thereby reducing the errors on the inferred distance. For the modeling of the longer evolution of Type II SN ejecta and their radiation, a time-dependent approach is warranted. To achieve a higher level of consistency, CMFGEN is under developement to follow as well the time-dependent evolution of the radiation field, together with options for chemical stratification and energy deposition from radioactive isotopes. This versatility will allow us to study SN ejecta of any type and over months after explosion." }, "0710/0710.2262_arXiv.txt": { "abstract": "We have developed a method for measuring higher-order weak lensing distortions of faint background galaxies, namely the weak gravitational flexion, by fully extending the Kaiser, Squires \\& Broadhurst method to include higher-order lensing image characteristics (HOLICs) introduced by Okura, Umetsu, \\& Futamase. We take into account explicitly the weight function in calculations of noisy shape moments and the effect of higher-order PSF anisotropy, as well as isotropic PSF smearing. Our HOLICs formalism allows accurate measurements of flexion from practical observational data in the presence of non-circular, anisotropic PSF. We test our method using mock observations of simulated galaxy images and actual, ground-based Subaru observations of the massive galaxy cluster A1689 ($z=0.183$). From the high-precision measurements of spin-1 first flexion, we obtain a high-resolution mass map in the central region of A1689. The reconstructed mass map shows a bimodal feature in the central $4'\\times 4'$ region of the cluster. The major, pronounced peak is associated with the brightest cluster galaxy and central cluster members, while the secondary mass peak is associated with a local concentration of bright galaxies. The refined, high-resolution mass map of A1689 demonstrates the power of the generalized weak lensing analysis techniques for quantitative and accurate measurements of the weak gravitational lensing signal. ", "introduction": "Propagation of light rays from a distant source to the observer is governed by the gravitational field of intervening mass fluctuations as well as by the global geometry of the universe. The images of background sources hence carry the imprint of the gravitational potential of intervening cosmic structures, and their statistical properties can be used to test the background cosmological models. Weak gravitational lensing is responsible for the weak shape-distortion and magnification of the images of background sources due to the gravitational field of intervening matter (e.g., Bartelmann \\& Schneider 2001; Umetsu, Futamase, \\& Tada 1999). To the first order, weak lensing gives rise to a few -- $10\\%$ levels of elliptical distortions in images of background sources, responsible for the second-order derivatives of the gravitational lensing potential. Thus, the weak lensing signal, measured from tiny but coherent quadrupole distortions in galaxy shapes, can provide a direct measure of the projected mass distribution of cosmic structures. However, practical weak lensing observations subject to the effects of atmospheric seeing, isotropic/anisotropic PSF, and (residual) camera distortion across the field of view, which must be examined from the stellar shape measurements and corrected for in the weak lensing analysis. Practical methods for PSF corrections and shear measurements/calibrations have been studied and developed by many authors, such as pioneering work of Kaiser, Squires, \\& Broadhurst (1995, hereafter KSB), the Shapelets technique which describes PSF and object images in terms of Gaussian-Hermite expansions (Refregier 2003), and recent systematic, collaborative efforts by The Shear TEsting Programme (Heymans et al. 2006; Massey et al. 2007). Thanks to these successful developments in weak lensing techniques as well as in instrument technology, the quadrupole weak lensing has become one of the most important tools in observational cosmology to map the mass distribution in individual clusters of galaxies (e.g., Kaiser \\& Squires 1993; Broadhurst et al. 2005a; Okabe \\& Umetsu 2007; Umetsu \\& Broadhurst 2007), measure ensemble-averaged mass profiles of galaxy-group sized halos from the galaxy-galaxy lensing signal (e.g., Hoekstra et al. 2001; Hoekstra et al. 2004; Parker et al. 2005; Mandelbaum et al. 2006), study the statistical properties of the large scale structure of the universe from the cosmic shear statistics (e.g., Bacon et al. 2000; van Waerbeke et al. 2001; Hamana et al. 2003), and search for galaxy clusters by their mass properties (e.g., Schneider 1996; Erben et al. 2000; Umetsu \\& Futamase 2000; Wittman et al. 2001). In recent years, there have been theoretical efforts to include the next higher order distortion effects as well as the usual quadrupole distortion effect in the weak lensing analysis (Goldberg \\& Natarajan 2002; Goldberg \\& Bacon 2005; Bacon et al. 2006; Irwin \\& Shmakova 2006; Goldberg \\& Leonard 2007; Okura, Umetsu, \\& Futamase 2007). We have proposed in Okura, Umetsu, \\& Futamase (2007, hereafter OUF) to use certain convenient combinations of octopole/higher multipole moments of background images which we call the Higher Order Lensing Image's Characteristics (HOLICs), and have shown that HOLICs serve as a direct measure for the next higher-order weak lensing effect, or the gravitational flexion (Goldberg \\& Bacon 2005) and that the use of HOLICs in addition to the quadrupole shape distortions can improve the accuracy and resolution of weak lensing mass reconstructions based on simulated observations. Recently, Goldberg \\& Leonard (2007) extended the HOLICs approach for flexion measurements to include observational effects, namely the Gaussian weighting in shape-moment calculations (see the appendix therein) and the isotropic PSF effect, under the assumption that PSF is nearly circular, and tested their extended HOLICs approach with simulated and HST/ACS observations. Leonard et al. (2007) have applied the extended HOLICs method to reconstruct the projected mass distribution in the central region of the massive galaxy cluster A1689 at $z=0.183$, and revealed substructures associated with small clumps of galaxies. Further Leonard et al. (2007) found that in dense systems such as galaxy clusters the HOLICs technique is robust and less sensitive than the Shapelet technique to contamination by light from the extended wings of lens/foreground galaxies. In the present paper we develop a method for measuring flexion by the HOLICs approach by fully extending the KSB formalism; We take into account explicitly the effects of Gaussian weighting in calculation of noisy shape moments and higher-order PSF anisotropy as well as isotropic PSF smearing. We then apply our method to actual, ground-based Subaru observations of A1689, and perform a mass reconstruction in the central region of A1689. The paper is organized as follows. We first summarize in \\S 2 the basis of weak gravitational lensing and the flexion formalism. In \\S 3, we derive the relationship between HOLICs and flexion by incorporating Gaussian smoothing in shape measurements in the presence of isotropic and anisotropic PSF. The practical method to correct the isotropic/anisotropic PSF effects will be presented in \\S 4. In Section 5 we use simulations to test our flexion analysis method based on the HOLICs moment approach. We then perform a weak lensing flexion analysis of A1689 by our fully-extended HOLICs approach, and perform a mass reconstruction of A1689 from the HOLICs estimates of flexion. Finally summary and discussions are given in \\S 6. We refer interested readers to a complete appendix\\footnote {Full appendix is available in electronic form at http://www.asiaa.sinica.edu.tw/keiichi/OUF2/appendix.pdf.} for details of the derivation of flexion-observable relationships in practical observations. ", "conclusions": "In the present paper, we have developed a method for weak lensing flexion analysis by fully extending the KSB method to include the measurement of HOLICs (OUF). In particular, we take into account explicitly the weight function in calculations of noisy shape moments and the effects of spin-1 and spin-3 PSF anisotropies, as well as isotropic PSF smearing, in the limit of weak lensing and small PSF anisotropy ($q$). The higher order weak lensing effect induces a centroid shift in the observed image of the background (Goldberg \\& Bacon 2005; OUF; Goldberg \\& Leonard 2007). In weighted moment calculations, this will yield in the flexion measurement additional correction terms (relevant to $W'(x), W''(x), W'''(x)$) that must be taken into account by properly expanding the weight function $W(x)$. It is found that neglecting these additional terms originated from the Taylor expansion of $W(x)$ yields the same result as obtained by Goldberg \\& Leonard (2007; see Appendix therein). We extended the KSB formalism to include the higher-order isotropic and anisotropic PSF effects relevant to spin-1 and spin-3 HOLICs by following the prescription given by KSB and Bartelmann \\& Schneider (2001), which provides direct relations between the observable HOLICs and underlying flexion in the weak lensing limit. We have implemented in our analysis pipeline our flexion analysis algorithm based on the HOLICs moment approach, and tested the reliability and limitation of our PSF correction scheme using numerical simulations. Our simulation results show that (i) after applying our PSF correction method the PSF-induced anisotropies in HOLICs of mock galaxy images can be considerably reduced by a factor of $10$--$100$, depending on the strength of PSF anisotropy, (ii) those small galaxies whose angular size is smaller than or comparable to the size of PSF suffer from severe anisotropic PSF effects, and that (iii) there is an overall trend that the fractional correction factor is larger for larger galaxy images. Therefore, our simulation results support the reliability of our PSF-correction scheme and its practical implementation. Based on the simulation results, we have applied our flexion analysis pipeline to ground-based $i'$ imaging data of the rich cluster A1689 ($z=0.183$) taken with Subaru/Suprime-Cam. Our flexion analysis of Subaru A1689 data revealed a non-negligible, significant effect of higher-order PSF anisotropy induced in stellar images (Figures \\ref{fig:zetaq_field} and \\ref{fig:zetaq}). It is therefore important in practical flexion measurements to quantify and correct for the higher-order anisotropic PSF effects. Our mass reconstruction from the first-flexion measurements shows two significant ($>4\\sigma$) mass structures associated with concentrations of bright galaxies in the central cluster region: the first peak ($5.2\\sigma$) associated with the central concentration of bright galaxies including the cD galaxy, and the second peak ($4.4\\sigma$) associated with a clump of bright galaxies located $\\sim 1'$ northeast of the cluster center. This significant detection of the second peak confirms earlier ACS results from the strong lensing analysis (Broadhurst et al. 2005b; Halkola et al. 2006; Leonard et al. 2007) and the combined strong lensing, weak shear, and flexion analysis by Leonard et al. (2007). The central mass peak, however, was not recovered in the earlier flexion analysis by Leonard et al. (2007) based on HST/ACS data. Leonard et al. (2007) attributed this to their relatively large reconstruction error at the cluster center, although they have a very large number density of background galaxies, $\\bar{n}_g\\approx 75$ arcmin$^{-2}$. On the other hand, owing to our conservative selection criteria for the background sample, the mean number density of background galaxies used for the present analysis is $\\bar{n}_g=7.75$ arcmin$^{-2}$, which is almost one order of magnitude smaller than that of the ACS data, and is about $20\\%-30\\%$ of a typical number density of magnitude/size-selected background galaxies usable for the quadrupole shape measurements in ground-based Subaru observations ($\\bar n_g\\sim 30-40 {\\rm arcmin}^{-2}$). However, we found that it is rather important to remove small/faint galaxy images and noisy outliers in flexion measurements since they are likely to be affected by the residual PSF anisotropy and/or observational noise in the shape measurement (\\S \\ref{subsec:sim}). Besides, the smaller the object, the larger the amplitude of intrinsic flexion contributions. Recall that flexion and HOLICs have a dimension of length inverse: The response to flexion is size-dependent, and the amplitude of intrinsic flexion is inversely proportional to the object size. Indeed, we find that inclusion of smaller objects results in a noisy reconstruction. Similar values of the background number density, $\\bar n_g\\simlt 10 {\\rm arcmin}^{-2}$, have been used in recent quadrupole weak lensing analyses based on Subaru observations (e.g., Broadhurst et al. 2005a; Umetsu \\& Broadhurst 2007; weak lensing cluster mass measurements of Okabe \\& Umetsu 2008). In their studies only objects redder than the cluster sequence are selected in color-magnitude space for their weak lensing analysis, because such a red population is expected to comprise only background galaxies ($\\bar z_s\\sim 0.9$; see, e.g., Medezinski et al. 2007), made redder by relatively large $k$-corrections and with negligible contamination by cluster galaxies (Broadhurst et al. 2005a; Medezinski et al. 2007). However, the smaller number of objects implies a coarser angular resolution in the map-making for achieving a proper signal-to-noise ratio (e.g., per-pixel ${\\rm S/N}\\simgt 1$). With $\\bar n_g\\sim 10 {\\rm arcmin}^{-2}$ for cluster quadrupole weak lensing, typical angular resolutions are about $1-2$ arcmin (e.g., Gaussian FWHM, or boxcar width). On the other hand, flexion measures essentially the gradient of the tidal gravitational shear field (i.e., $F,G\\propto \\phi(r)/r^3$), and hence is relatively sensitive to small-scale structures. Therefore, our successful reconstruction of the mass substructures with a small background density, $\\bar n_g\\sim 8 {\\rm arcmin}^{-2}$, could be attributed to the superior sensitivity of flexion to small scale structures (see OUF for detailed discussions) and the here-adopted selection criteria for a background galaxy sample for weak lensing flexion analysis. Finally, we emphasize that our HOLICs formalism here is different from the earlier work by Goldberg \\& Leonard (2007) in that (1) additional correction terms for the centroid shift, relevant to the derivatives of the weight function, have been included and (2) the spin-1 and spin-3 PSF anisotropies, as well as the isotropic PSF smearing, have been taken into account under the assumption of small PSF anisotropy ($q[\\btheta]$), as done in the KSB formalism. Our flexion-based mass reconstruction of A1689 demonstrates the power of the generalized flexion analysis techniques for quantitative and accurate measurements of the weak gravitational lensing effects." }, "0710/0710.5383_arXiv.txt": { "abstract": "We present further \\sirtf\\ Space Telescope observations of the recurrent nova {\\rs}iuchi, obtained over the period 208--430~days after the 2006 eruption. The later \\sirtf\\ IRS data show that the line emission and free-free continuum emission reported earlier is declining, revealing incontrovertible evidence for the presence of silicate emission features at 9.7 and 18\\mic. We conclude that the silicate dust survives the hard radiation impulse and shock blast wave from the eruption. The existence of the extant dust may have significant implications for understanding the propagation of shocks through the red giant wind and likely wind geometry. ", "introduction": "\\rs\\ is a recurrent nova (RN) that erupted in 1898, 1933, 1958, 1967, 1985 \\citep{wallerstein}. It consists of a semi-detached binary (orbital period 455.7~days) comprising a roche-lobe-filling red giant (M2III) and a massive ($\\gtsimeq1.2\\Msun$) white dwarf \\citep[WD;][]{fekel}. The eruption follows a thermonuclear runaway on the surface of the WD \\citep{tnr}. In the case of the \\rs\\ class of RN, however, the ejected material runs into, and shocks, a dense red giant (RG) wind \\citep{bk85,obrien92}. The 1985 eruption was, for the first time, the subject of a multi-wavelength observational campaign, from the radio to the X-ray \\citep{vnu}. Its most recent eruption, on 2006 February 12.83 \\citep[][ we take this to define the origin of time post-outburst]{hirosawa}, was the subject of an even more intensive observational campaign. Infrared (IR) observations \\citep{das06,evans07a,evans07b} showed evidence for the shock, seen also at radio \\citep{eyres,obrien} and X-ray \\citep{bode06,sokoloski,ness,osborne} wavelengths, as the ejecta interacted with the RG wind. Observations of \\rs\\ obtained with the Spitzer Space Telescope \\citep[{\\it Spitzer};][]{werner, gehrz07} during the period from 2006 April 16 through 26~UT ($\\simeq 67$ days after outburst) were described by \\cite{evans07b}. We present here further IR observations conducted as part of a long duration synoptic campaign, obtained later in the outburst with the \\sirtf\\ Infrared Spectrometer \\citep[IRS;][]{houck}. ", "conclusions": "We have presented further \\sirtf\\ IRS observations of \\rs\\ in the aftermath of its 2006 eruption. Unlike our earlier observation, which showed emission by the hot shocked gas, more recent IR spectra reveal the presence of silicate dust. The dust we see could not have been formed in the 2006 eruption. It most probably survived the 2006 eruption, because a portion of the RG wind (including the dust) did not see the eruption at all, and because the passing shock failed to destroy any dust that survived the UV blast at the eruption. This conclusion may have major implications for the evolution of the shock, and of the long-term survival of the RG wind between eruptions. We predict that the dust we see around \\rs\\ will persist until the next eruption." }, "0710/0710.0790_arXiv.txt": { "abstract": "{Water vapor emission at 22 GHz from masers associated with star-forming regions is highly variable.}{We present a database of up to 20 years of monitoring of a sample of 43 masers within star-forming regions. The sample covers a large range of luminosities of the associated IRAS source and is representative of the entire population of \\hdo\\ masers of this type. The database forms a good starting point for any further study of \\hdo\\ maser variability.}{The observations were obtained with the Medicina 32--m radiotelescope, at a rate of 4--5 observations per year.} {To provide a database that can be easily accessed through the web, we give for each source: plots of the calibrated spectra, the velocity--time--flux density plot, the light curve of the integrated flux, the lower and upper envelopes of the maser emission, the mean spectrum, and the rate of the maser occurrence as a function of velocity. Figures for just one source are given in the text for representative purposes. Figures for all the sources are given in electronic form in the on-line appendix. A discussion of the main properties of the \\hdo\\ variability in our sample will be presented in a forthcoming paper.} {} ", "introduction": "Since the discovery of 22 GHz \\hdo\\ maser emission associated with young stellar objects (YSOs) within star-forming regions (SFRs), variability of maser emission is well-known. Changes as large as several orders of magnitude in the maser emission have been observed (e.g., Little et al.~\\cite{lit77}; Liljestr\\\"om et al.~\\cite{lil89}; Claussen et al.~\\cite{cla96}; Comoretto et al.~\\cite{com90}; Wouterloot et al.~\\cite{wou95}). At the same time, velocity drifts of individual components of up to a few \\kms\\ per year have also been reported (e.g., Brand et al.~\\cite{BCCFPPV2003}). The variability can be slow or burst-like and covers all ranges of timescales, from hours-days to months-years. In the present study we are mainly concerned with the latter and, for the first time, deal with a large sample of sources (43) and with a time-span of up to 20 years, with 4--5 spectra per year. Due to the large amount of telescope time required to follow the evolution of \\hdo\\ maser emission, inevitably the variability is more easily monitored through single dish observations. In fact, besides our campaign only the Pushchino 22--m single-dish maser patrol covers a comparably long period (e.g., Rudnitskij et al.~\\cite{russi}). Furuya et al.\\ (\\cite{furuya}) carried out a multiepoch 22 GHz H$_{2}$O maser survey towards 173 low-mass YSOs (Class 0 to Class III sources) using the Nobeyama 45 m telescope. This was the first complete water maser survey towards Class 0 sources in the northern sky. However, their observations extend non-uniformly over a period of only three years. It is well-known that the variability depends strongly on the luminosity of the SFR (usually derived from the associated IRAS source). For instance, there is a minimum luminosity ($\\sim$25 L$_\\odot$) below which \\hdo\\ maser emission may be present only for about one third of the entire duration of the maser activity (Persi et al.~\\cite{per94}; Claussen et al.~\\cite{cla96}). The results obtained so far suggest that the lower the SFR luminosity, the higher the observed degree of variability of the \\hdo\\ maser emission. Higher luminosity sources may show more steady components. Consequently, a large sample of sources is needed to better understand the dependence of the variability on other parameters of the SFRs, in particular on their luminosities which cover a range of several orders of magnitude, and on their (molecular) environment. The use of a single dish instrument is of course a limitation because interferometric observations clearly show that \\hdo\\ masers within a SFR often consist of many spatially separated, unresolved components, generally clustered in groups (usually with different velocities), which in most cases cannot be separated in single dish observations (e.g., Forster \\& Caswell~\\cite{for89}, ~\\cite{for99}; Tofani et al.~\\cite{tof95}). However, Felli et al.~(\\cite{FMRC2006}) reported a case in which single dish observations were able to separate the evolution of spatially distinct components. Nevertheless, frequent single dish observations have the advantage of being able to follow potential velocity drifts of individual {\\it velocity} components, which can be easily identified in the spectra, and thus give a better view of the dynamics (e.g., accelerations) occurring in the circumstellar environment of the YSO. In two earlier papers (Valdettaro et al.~\\cite{VPBCCFP2002}; Brand et al.~\\cite{BCCFPPV2003}), we set the basis for a systematic study of the \\hdo\\ variability in SFRs with the Medicina 32--m radiotelescope using a sample of 14 sources which covered a wide range of luminosities and had been observed for more than 10 years. Here we report on a larger sample of 43 SFRs (including the former 14 sources) and increase the time coverage up to 20 years. The database is presented in the form of plots of the calibrated spectra and additional plots of derived quantities, in a form that can be easily accessed through the web. In a forthcoming paper, we will analyze the properties of \\hdo\\ variability of our sample. ", "conclusions": "We present the observational results of a systematic study extending for almost 20 years of the \\hdo\\ maser variability in 43 SFRs with luminosities of the associated IR sources between 20 and 1.5$\\times 10^6$ L$_\\odot$. This database provides the backbone for the discussion of the main long-term properties of maser emission that will be presented in a forthcoming paper. We have identified several ways to describe graphically the main aspects of the \\hdo\\ maser emission. These include: \\begin{itemize} \\item [(a)] the spectra in a compressed form, with an autoscaled flux density scales (the same sets of plots, but with fixed linear and logarithmic flux density scales, are also available from our WEB pages); \\item [(b)] the velocity--time--flux density plots, which conveniently describe the morphology of the variability of the maser emission. These diagrams are particularly useful for recognizing the presence of possible velocity drifts and separating steady components from bursts of short duration; \\item [(c)] the velocity--integrated flux density as a function of time. This light curve describes the overall emission of the maser components associated with a SFR; \\item [(d)] the mean spectrum; \\item [(e)] the upper and lower envelopes of the maser emission. The upper envelope represents the maximum emission that the SFR could produce if {\\it all} the velocity components were present {\\it simultaneously} and emitting at their {\\it maximum} rate. Similarly, the lower envelope pinpoints the steady components and their lowest level (but $> 5 \\sigma$) of emission; \\item [(f)] the number of times that the maser emission is well above the noise ($> 5 \\sigma$) as a function of velocity is directly gauged by the histogram of the rate--of--occurrence. \\end{itemize}" }, "0710/0710.0759_arXiv.txt": { "abstract": "{ We suggest to use the observationally measured and theoretically justified correlation between size and rotational velocity of galactic discs as a viable method to select a set of high redshift standard rods which may be used to explore the dark energy content of the universe via the classical angular-diameter test. Here we explore a new strategy for an optimal implementation of this test. We propose to use the rotation speed of high redshift galaxies as a standard size indicator and show how high resolution multi-object spectroscopy and ACS/HST high quality spatial images, may be combined to measure the amplitude of the dark energy density parameter $\\Omega_{Q}$, or to constrain the cosmic equation of state parameter for a smooth dark energy component ($w=p/\\rho, \\;\\; -1\\le w < -1/3$). Nearly 1300 standard rods with high velocity rotation in the bin $V=200\\pm 20$km/s are expected in a field of 1 sq. degree and over the redshift baseline $0$12 days and short outbursts lasting $<$12 days. They assigned the letter \"L\" to the long outbursts and \"S\" to the short and anomalous outbursts. The anomalous outbursts tend to have a linear, lower rate of change and smaller amplitude than long and short outbursts (Figure 1). They found the most common sequence as LS (with 134 occurrences), LLS (69), LSSS (14), and LLSS (8). About 89\\% of all outbursts can fall into one of these four sequences. \\begin{figure} \\epsscale{1} \\plotone{f1.eps} \\caption{A sample of the SS Cyg visual light curve showing the three category of outbursts: long, short and anomalous (left to right).} \\end{figure} \\subsection{Outburst Cycles \\& Periodicity} Honey et al. (1989) photometrically observed SS Cyg over a single long and a single short outburst cycle. They reported substantial flickering on the order of minutes but were unable to find any modulation (including near the orbital period) or periodicity to an amplitude limit of 0.05 magnitude. Giovannelli, Martinez-Pais and Graziati (1992) review the flickering behavior of SS Cyg in both outburst and quiescence, but find no correlation with outburst type, although they do find an inverse correlation between flickering amplitude and system brightness. Cannizzo and Mattei (1992) analyzed the historical light curve of the American Association of Variable Star Observer's (AAVSO) International Database (hereafter: AID), which at the time included 29,387 individual daily means from 1896 September 27 to 1992 April 7. They found no correlation between outburst duration time and cycle time but did confirm a correlation between quiescent magnitude and cycle time, caused by a variation in the mass transfer rate of up to a factor of 2 on yearly time scales and a variation of 20\\%-30\\% on decadal time scales. They did not find any periodicity in cycle time. In a later study \\citep{can98}, using a smaller subset of the data (1963 - 1997), they discovered a variable decay rate in some outburst declines. In a minority of the outbursts, the decay rate slows down about two-thirds of the way into the decline for about a day and a half, then the decay returns to its previous rate. The significance of the break is proportional to the length of the total decline. The break manifests itself as a 20\\%-300\\% decrease in the decay rate for around one day. Cannizzo and Mattei (1998) refers to this phenomenon as a \"glitch\", we refer to it hereafter as the \"Cannizzo Glitch\". Ak et al. (2001) analyzed the quiescence magnitude and outburst cycles of 23 dwarf novae mostly using Fourier analysis. They found no correlation between outburst cycle period and masses of component stars, mean outburst interval, mean outburst duration, mean decline and rise rates of outbursts, absolute quiescent magnitudes and outburst states. In a poster, Hill and Waagen (2005) suggested the existence of a pre-outburst brightening of about 2\\% over the course of around 12 hours followed by a plateau which lasts another 12 hours leading up to the onset of the outburst. The analyzed data set included about half of the AID visual data and excluded CCD observations. This was the only search for predictors found in published literature. \\subsection{Long Term Variation} Kiplinger, et al. (1988) report a 0.15 magnitude amplitude, 7.2 year period in the quiescent behavior of SS Cyg, determined through Fourier analysis of one day means in the AID data from 1896 - 1984. Also, Bianchini (1988) discovered a similar 6.9 year period through Fourier analysis of the intervals between outbursts and attributes it to solar type variation in the secondary \\citep{Bia92}. However, Richman, Applegate and Patterson's (1994) own analysis of the individual data points plus various long term moving averages find that the power of the reported $\\sim$7 year period is not significantly greater than that of other low frequency signals in the power spectrum and caution that the reported significance is due to the eye being, \"...very prone to spot one to three cycles of periodic behavior in any randomly varying time series.\" However, their analysis excluded all quiescent data points within 12 days on either side of the outburst, thus if the source for the reported $\\sim$7 year period was found in activity near the outburst then it would be hidden from their analysis. Hemplemann and Kurths (1990) O-C analysis found secular variation within 100 years as well as deviations from the sample mean over intervals of a few tens of cycles. Jevtic, Mattei and Schweitzer (2003) performed a nonlinear analysis using Poincar\\'{e} section and found a period of around 52 years. It is unclear whether the two periods are related as a fundamental and a harmonic pair. ", "conclusions": "We have analysed intensive {\\it V} and {\\it I$_{c}$} band observations of SS Cyg over two observing seasons. We were unable to detect any predictors or other activity which could be used to predict an oncoming outburst. However, the combined uncertainty of our photometric data was high enough to warrant further investigation with more precise observations. In particular, some of the data suggest a rise in the ({\\it V-I$_{c}$}) color beginning five days prior to outburst. The rise level was barely within our uncertainty so we cannot report it. A similar dataset in ({\\it B-$I_{c}$}) may increase the photometric sensitivity to such a feature, if it exists. We also analysed 102 years of AAVSO visual observations. No periodicity was detected which had not already been reported. However, long term quasiperiodic features were detected on the order of 1,000-2,000 days. Their existence is moderately correlated with shortened intervals between outbursts. If the quasiperiodic features are due to solar type variation, as previously reported, then it is possible that the enhanced activity drives additional mass transfer thus decreasing the mean time between outbursts." }, "0710/0710.1872_arXiv.txt": { "abstract": "We present a new \\sch orbit-superposition code that is designed to model discrete datasets composed of velocity measurements of individual kinematic tracers in a dynamical system. This constitutes an extension of previous implementations that can only address continuous data in the form of (the moments of) velocity distributions, thus avoiding potentially important losses of information due to data binning. Furthermore, the code can handle any combination of available velocity components, i.e., only line-of-sight velocities, only proper motions, or a combination of both. It can also handle a combination of discrete and continuous data. The code determines the combination of orbital mass weights (representing the distribution function) as a function of the three integrals of motion $E,L_z,$ and $I_3$ that best reproduces, in a maximum-likelihood sense, the available kinematic and photometric observations in a given axisymmetric gravitational potential. The overall best fit is the one that maximizes the likelihood over a parameterized set of trial potentials. The fully numerical approach ensures considerable freedom on the form of the distribution function $f(E,L_z,I_3)$. This allows a very general modeling of the orbital structure, thus avoiding restrictive assumptions about the degree of (an)isotropy of the orbits. We describe the implementation of the discrete code and present a series of tests of its performance based on the modeling of simulated (i.e., artificial) datasets generated from a known distribution function. We explore pseudo-datasets with varying degrees of overall rotation and different inclinations on the plane of the sky, and study the results as a function of relevant observational variables such as the size of the dataset and the type of velocity information available. We find that the discrete \\sch code recovers the original orbital structure, mass-to-light ratio, and inclination of the input datasets to satisfactory accuracy, as quantified by various statistics. The code will be valuable, e.g., for modeling stellar motions in Galactic globular clusters, and modeling the motions of individual stars, planetary nebulae, or globular clusters in nearby galaxies. This can shed new light on the total mass distributions of these systems, with central black holes and dark matter halos being of particular interest. ", "introduction": "\\label{sec.intro} The study of the internal dynamics of stellar systems plays an essential role in astronomy. From the observed positions and velocities of the stars in galaxies and globular clusters it is possible to infer their total (dark+luminous) mass distribution, which, in particular, provides information on the presence and properties of dark halos and massive black holes. In turn, this structural knowledge constrains theories for the formation and evolution of these systems. The dynamical state of a stellar system is determined by its phase space distribution function, $f({\\vec r}, {\\vec v})$, which counts the stars as a function of position ${\\vec r}$ and velocity ${\\vec v}$. Typically, however, only three of the six phase-space coordinates are available observationally: the projected sky position $(x',y')$, and the velocity $v_{z'}$ along the line of sight (LOS). Proper motion observations can provide the additional velocities $(v_{x'},v_{y'})$, but such data are generally not available (with the notable exception of some Galactic globular clusters). To make progress with the limited information available, the dynamical theorist is often forced to make simplifying assumptions about geometry (e.g., that the system is spherical) or about the velocity distribution (e.g., that it is isotropic). Such assumptions can have strong effects on the inferred mass distribution (\\citealt{bin82}). To obtain the most accurate results it is therefore important to make models that are as general as possible. Of particular importance for collisionless, unrelaxed systems such as galaxies is to constrain the velocity anisotropy using available data, rather than to assume it a priori. In a collisionless system the distribution function satisfies the collisionless Boltzmann equation. Analytical methods to find solutions of this equation usually rely on the Jeans Theorem, which states that the distribution function must depend on the phase-space coordinates through integrals of motion (quantities that are conserved along a stellar orbit). In a spherical system all integrals are known analytically, namely, the energy $E$ and the components of the angular momentum vector ${\\vec L}$. Analytical models for spherical systems are therefore fairly easily constructed. In an axisymmetric system things are more complicated (e.g., \\citealt{bt87,mer99}). Only two integrals are known analytically, $E$ and the vertical component $L_{\\rm z}$ of the angular momentum vector\\footnote{We adopt the notation in which $(x,y,z)$ denote the coordinates intrinsic to the axisymmetric stellar system, with the plane $x-y$ being the equatorial plane, and $z$ the symmetry axis. These relate via the inclination $i$ to the observable coordinates $(x',y')$ on the plane of the sky (aligned, respectively, along the projected major and minor axes of the stellar system), and $z'$ the line-of-sight direction, positive in the direction away from us.}, but there is generally a third integral for which no analytical expression exists. Therefore, it is not generally possible to construct an axisymmetric model analytically. The special class of so-called `two-integral' ($f=f(E,L_z)$) models (e.g., \\citealt{bat93,deh94,ver02}) has its uses (e.g., \\citealt{mag98,vdm06}), but these have an isotropic velocity distribution in their meridional plane, which need not be a good fit to real dynamical systems. The most practical way to model a general axisymmetric system is to do it numerically. While a few methods exist to do this (e.g., \\citealt{m2m,nmagic}), the most common approach uses Schwarzschild's (1979) method. One starts with a trial guess for the gravitational potential $\\Psi$ and then numerically calculates an orbit library that samples integral space in some complete and uniform way. The orbits are integrated for several hundred orbital periods, and the time-averaged intrinsic and projected properties (density, LOS velocity, etc.) are stored as the integration progresses. The construction of a model consists of finding a weighted superposition of the orbits that: (1) reproduces the observed stellar or surface brightness distribution on the sky; and (2) reproduces all available kinematical data to within the observational error bars. Additional constraints can be added to enforce that the distribution function in phase space be smooth and reasonably well behaved, e.g., through regularization or by requiring maximum entropy. Several axisymmetric Schwarzschild codes have been developed in the last decade (e.g., \\citealt{vdm98,cre99,geb00,val04,tho04}). These codes deal with the situation in which information on the line-of-sight velocity distribution (LOSVD) is available for a set of positions on the projected plane of the sky. This is the case, e.g., when the kinematical data are from long-slit or integral-field spectroscopic observations of unresolved galaxies. The optimization problem for such data can be reduced to a linear matrix equation for which one needs to find the least-squares solution with non-negative weights \\citep{rix97}. One dimension of the matrix corresponds to the number of orbits in the library, while the other corresponds to the number of (luminosity, kinematical and regularization) constraints that must be reproduced. Both dimensions are typically in the range $10^3$--$10^4$. Nonetheless, efficient numerical algorithms exist to find the solution, which yield the orbital and the velocity distribution of the model, as well as the $\\chi^2$ of the fit to the kinematical data. The procedure must then be iterated with different gravitational potentials, to determine the potential that provides the overall best $\\chi^2$. The existing codes have been used and tested extensively (e.g., \\citealt{cre00,cap02,cap06,geb03,ben05,dav06}). Some questions remain, e.g., about the importance of smoothing in phase space, the exact meaning of the confidence regions determined using $\\Delta \\chi^2$ contours, and, in some situations, valid concerns have been raised regarding whether the available data contain enough information so as to warrant the conclusions of the \\sch modeling \\citep{val04,cre04,kra05}. Nevertheless, on the whole Schwarzschild codes have now been established as an accurate and versatile tool to study a wide range of dynamical problems. A disadvantage of the existing codes is that they cannot be easily applied to the large class of problems in which the kinematical observations come in the form of discrete velocity measurements, rather than as LOSVDs. This is encountered, e.g., when modeling the dynamics of galaxies at large radii, where the low-surface brightness prevents integrated-light spectroscopy. The only available data are then often of a discrete nature, e.g., via the LOS velocities of individual stars in galaxies of the Local Group (e.g., \\citealt{bill00,jan01,jan02,lok02,wil04,lok05,wal06,geh06}), or via planetary nebulae (e.g., \\citealt{dou02,rom03,teo05}) and globular clusters (e.g., \\citealt{cote01,tom04}) surrounding giant ellipticals. The kinematical data available for clusters of galaxies, consisting of redshifts for individual galaxies, are of a similarly discrete nature (e.g., \\citealt{lok03}). The typical datasets in all these cases consist of tens to hundreds of LOS velocities. Galactic globular clusters constitute another class of object for which kinematical data is often available only as discrete measurements, rather than in the form of LOSVDs. From ground-based observations, data sets of individual LOS velocities can be available for up to thousands of stars in these systems (e.g., \\citealt{sun96,may97,rei06}), and for $\\omega$ Cen it has been possible to assemble large samples of proper motions as well \\citep{vleu00}. With the capabilities of {\\it HST}, accurate proper motion data sets with up to $\\sim 10^4$ stars are now becoming available for several more Galactic globular clusters (e.g., \\citealt{mcn03,mcl06}). Note that discrete datasets do not necessarily provide better or worse information than datasets obtained from integrated-light measurements. Both types of data have their advantages and disadvantages. For discrete datasets, for example, interloper contamination can be a problem (see also the end of Section~\\ref{sec:logL} below). By contrast, for integrated-light measurements, it is often difficult to constrain the wings of the LOSVD due to uncertainties associated with continuum subtraction. Which type of data is most appropriate and most easily obtained depends on the specific object under study. This is therefore not a question that we address in this paper. Instead, we focus on the issue of how to best analyze discrete data, if that happens to be what is available. Analyses of discrete datasets have often been more simplified than the analyses that are now common for integrated-light data. For example, the observations are analyzed using the Jeans equations (e.g., \\citealt{ger02,lok03,cote03,dou07}), often with the help of data binning to calculate rotation velocity and velocity dispersion profiles (see, however, the ``spherical'' Schwarzschild models of M87 of \\citealt{rom01}). The disadvantage of such an approach is that not all the information content of the data is used, including information on deviations of the velocity histograms from a Gaussian. Such deviations are important because they constrain the velocity dispersion anisotropy of the system (e.g., \\citealt{vdm93,ger93,ger98}). This anisotropy is an important ingredient in some existing controversies, e.g. regarding the presence of dark halos around elliptical galaxies \\citep{rom03,dek05}. Loss of information can be avoided when large numbers of datapoints are available, as is often the case for globular clusters. It is then possible to create velocity histograms for binned areas on the projected plane of the sky, after which analysis can be done with existing Schwarzschild codes (e.g., \\citealt{bos06}). While this is possible for large datasets, such an approach is not viable for the more typical, smaller datasets that are often available. The availability of Schwarzschild codes that can fully exploit the information content of such smaller datasets would therefore be valuable to advance this subject. Motivated by these considerations we set out to adapt our existing Schwarzschild code \\citep{vdm98} to deal with discrete datasets. This does not constitute a trivial change, since it changes the constrained superposition procedure from a linear matrix problem to a more complicated maximum likelihood one. For each observed velocity of a particle in the system the question becomes: what is the probability that this velocity would have been observed if the model is correct? The overall likelihood of the data, given a trial model, is the product of these probabilities for all observations. Such likelihood problems have previously been solved for spherical systems \\citep{mer93,vdm00,wu06} and the special class of axisymmetric $f(E,L_z)$ systems \\citep{mer97,wu07}. However, for the axisymmetric Schwarzschild modeling approach the problem corresponds to finding the minimum of a function in a space with a dimension of $10^3$--$10^4$. We show in this work, via the \\sch modeling of simulated datasets, that this problem can indeed be solved successfully and efficiently. Moreover, we follow \\cite{glenn06} and implement in our new code the ability to calculate and fit proper motions in addition to LOS velocities. Applications of the code to real datasets will be presented in forthcoming papers. The structure of the paper is as follows. In Section \\ref{sec:logL} we phrase the new problem of fitting a \\sch model to a dataset of discrete velocities (of one, two, or three dimensions) of individual kinematic tracers in terms of a likelihood formalism. Section \\ref{sec:code} describes the implementation of the discrete fitting procedure into our existing \\sch code. At the same time, we summarize here the major steps involved in the construction of the probability matrix that describes the likelihood of a given kinematic data point belonging to some particular orbit of the library. We then present in Section \\ref{sec:tools} sets of simulated data that we use for the purpose of testing the performance of the discrete \\sch code. We also describe the known input distribution functions from which these data were drawn. The application of the code to the simulated datasets is presented in Section \\ref{sec:tests}. We present a thorough analysis of the accuracy with which our discrete \\sch code recovers the known distribution function, mass-to-light ratio and inclination used to generate the simulated data. Finally, in Section \\ref{sec:end} we summarize our findings and present our conclusions. ", "conclusions": "\\label{sec:end} Discrete kinematic datasets, composed of velocities of individual tracers (e.g., red giants, planetary nebulae, globular clusters, galaxies, etc.), are routinely being assembled for a variety of stellar systems of all scales (\\S\\,\\ref{sec.intro}). These include not only LOS-velocity surveys. High-quality proper-motion databases already exist for Galactic globular clusters, and future facilities hold the promise of providing the same for stars in the nearest galaxies. However, the most sophisticated tools typically being used in the modeling of these observations were actually developed for the analysis of kinematic data in the form of LOSVDs, a rather different type of velocity information than the case of the velocities of kinematic tracers on a one-by-one basis. As a consequence, the information content of any particular dataset of a discrete nature is likely not being fully exploited. We thus have developed a specific tool for the modeling of discrete datasets, which we have presented in this paper along with detailed tests of its performance based on the modeling of simulated data. The new tool consists of a \\sch orbit-superposition code that, adapted from the implementation of \\citet{vdm98}, can handle any number of (one-, two-, or three-dimensional) velocities of individual kinematic tracers without relying on any binning of the data. Under the only assumptions that the system is in steady-state equilibrium (i.e., the gravitational potential is not changing in time) and may be well approximated as axisymmetric, the code finds the distribution function (a function of the three integrals of motion $E$, $L_z$, and $I_3$) that best reproduces the observations (the velocities of the tracers as well as the overall light distribution) in a given potential. The fact that the distribution function is free to have any dependence on the three integrals of motion allows for a very general description of the orbital structure, thus avoiding common restrictive assumptions about the degree of (an)isotropy of the orbits. Unlike previous implementations of the \\sch technique, we cast the problem of finding the best superposition of orbits using a probabilistic approach, i.e., by building a likelihood function representing the probability that the entire set of measurements would have been observed assuming a particular form for the gravitational potential (\\S\\,\\ref{sec:logL}). In this case, and in contrast with the old continuous versions, the dependence of the likelihood function on the orbital weights is non-linear, and the optimization problem can not be reduced to a linear matrix equation. Instead, it becomes a problem of the maximization of a likelihood with respect to the set of weights associated to all possible combinations of the integrals $(E,L_z,I_3)$ that comprise the orbit library (\\S\\,\\ref{sec:logL}), and which accounts for the observed positions and (any-dimensional) velocities of all particles in the dataset, including their uncertainties (\\S\\,\\ref{sec:pij}). After extensive testing, a conjugate gradient algorithm was found to converge satisfactorily to the correct solution and was adopted for the remaining tests of the code's overall performance (\\S\\,\\ref{sec.mkfitin}). In order to assess the reliability of our discrete \\sch code, we applied it to several sets of simulated data, i.e., artificially generated kinematic observations obtained from a model of an axisymmetric galaxy of which the orbital structure, mass distribution, and inclination are known in advance. Pseudo-datasets were generated from a two-integral phase-space distribution function with varying degrees of overall rotation, types of velocity information (only-LOS, only proper motions, and both), total number of particles, and for two different inclinations on the plane of the sky (\\S\\,\\ref{sec:data}). Using the various simulated datasets, we studied the recovery of the input orbital structure or DF, mass-to-light ratio, and inclination. For the purposes of these tests, we assumed complete knowledge of the radial profile of the underlying mass distribution and a mass-to-light ratio that remains constant as a function of radius. These restrictions are easily (and must be) lifted when modeling data on real systems, in which case one needs to explore a range of plausible underlying potentials and allow for variations of the mass-to-light ratio to properly account for the possibility of central black holes and dark halos. Inside the region constrained by data, we find that the distribution function (represented by the corresponding distributions of orbital mass weights) and streaming characteristics of the input datasets are satisfactorily recovered by the \\sch fits when the correct inclination and mass-to-light ratio are known (Figs. \\ref{fig:1Dplots} to \\ref{fig:Ebins55is}). As measured by the mean absolute deviations between the integrated weight distributions, the agreement between the fitted and the input orbital weight distributions as a function of $E$, $L_z$, and $I_3$ is typically of the order of 3\\%, 10\\%, and 20\\%, respectively (the numbers for our worst case being 5\\%, 16\\%, and 25\\%). When eliminating the dependence on $I_3$, the agreement between the fitted and input $E-L_z$ distributions is of the order of 15\\%, with the net rotational behavior of the input datasets cleanly recovered (Figs. \\ref{fig:2Dplot55ns} and \\ref{fig:2Dplot55is}). Thus, we conclude that the discrete \\sch code can successfully recover the orbital structure of the system under study. Assuming that the inclination of the system on the plane of the sky is known, we quantified the recovery of the input mass-to-light ratio as a function of the size of the input dataset (Fig. \\ref{fig:MLparabN}) and of the type of kinematic information available (Fig. \\ref{fig:MLparab2}). We studied both the best-fit value as well as the uncertainty in its determination (Fig. \\ref{fig:errors_ML}). The statistical expectation of better results when the amount of observational information is larger (either regarding the number of datapoints or the number of velocity components) is clearly reproduced by our discrete \\sch models. For the smallest datasets used in our testing ($N=100$), and regardless of whether using only-LOS velocities, only proper motions, or both, the best-fit mass-to-light ratio is within 5-10\\% of the input value, with formal $1\\sigma$ uncertainties of the order of 15\\%. When increasing either the number of available measurements or the number of measured velocity components, the mass-to-light ratio is always recovered to better than $\\sim 10\\%$ accuracy, with the corresponding random ($1\\sigma$) uncertainties in the range of 5-10\\%. The discrete \\sch code, therefore, recovers the mass-to-light ratio of the input datasets to satisfactory levels of accuracy. The recovery of both the mass-to-light ratio and inclination when neither of these quantities are known in advance (as is usually the case with real observations) was studied using a grid of discrete \\sch models, exploring also the dependence on the type of velocity components available (Fig. \\ref{fig:55is_LOS_MU}). We find that the mass-to-light ratio was again successfully recovered, but the best-fit inclination was not identified correctly using small orbit libraries. We found that this was remedied by better sampling the available $(E,L_z,I_3)$ integral space using a larger orbit library (Fig. \\ref{fig:grid_55_90}). For our input datasets with $i=55\\grad$, the best-fit inclination obtained by our models with a large orbit library is $57\\grad$, while for input datasets with $i=90\\grad$ we obtain a best-fit model with $i=80\\grad$. Given the known difficulty of \\sch models in general for determining the inclination of stellar systems, and considering the low relative importance of this parameter compared to other properties such as the orbital structure and the mass-to-light ratio, we regard this small disagreement for the high inclination datasets as acceptable. In summary, we have shown that our new \\sch code, designed to adequately handle modern datasets composed of discrete measurements of kinematic tracers, doing this without any loss of information due to data binning or restrictive assumptions on the distribution function, is able to constrain satisfactorily the orbital structure, mass-to-light ratio, and inclination of the system under study. Applications to data for Galactic globular clusters and nearby dE galaxies will be presented in future papers. These are only two examples of a large range of dynamical problems in astronomy to which a discrete \\sch code like ours can be applied, so we expect this new tool will contribute to the better understanding of stellar systems in general." }, "0710/0710.1043_arXiv.txt": { "abstract": " ", "introduction": "The detection of broad polarised lines in Seyfert~2 nuclei stands that this type of AGNs are intrinsically the same objects than Seyfert~1 galaxies and the differences in their observational properties can be understood as orientation effects. Within the Unified Models, the central continuum source of AGNs and the broad line region (BLR) are surrounded by an optically and geometrically thick structure of dust and molecular gas, likely following a toroidal geometry. The orientation with respect to our line of sight of Seyfert~2 galaxies is such that emission from the central engine is shielded by the molecular torus and therefore its continuum and broad emission lines are only observed in polarised light as a result of scattering outside the torus. Firstly confirmed in NGC~1068 (Miller et al. 1991), the central emission in Seyfert~2 is scattered by free electrons in a conical structure placed along the axis of the torus, the so-call ionisation cones, generating a polar scattering spectrum with a position angle perpendicular to the scattering cone. In contrast, the polarised spectrum of Seyfert~1 galaxies does not conform with simple polar scattering, exhibiting a wide diversity of polarisation properties. These results state that the geometry of a single polar scattering region is incomplete to explain simultaneously all types of Seyfert polarised spectra. Based on a study of 36 Seyfert~1 galaxies Smith et al. (2002) proposed a model in which the broad-line emission is originated in a rotating disk and scattered in two different regions: the classical {\\it ionisation cones} responsible of the polar scattering and an equatorial scattering region located within the torus and co-planar with the rotating disc. Therefore, the two extreme cases are the face--on Seyfert~1 which exhibit a null or weak polarisation as a result of cancellation of polar and equatorial scattering and the Seyfert~2, dominated by polar scattering as the equatorial scattering region is completely obscured. In Seyfert~1 with an intermediate line of sight angle, both scattering regions are visible but in general equatorial polarisation dominates. Interestingly, a peculiar type of Seyfert~1 galaxies exhibits polarised spectra similar to those of Seyfert~2, i.e. dominated by polar scattering. According to the model proposed by Smith et al. (2002), these {\\it polar--scattered} Seyfert 1 galaxies should be observed at an inclination comparable with the torus opening angle, and suffer therefore only a moderate extinction through the torus rim. Smith et al. (2004) estimate that between 10\\% and 30\\% of Seyfert~1 galaxies are dominated by polar scattering. Their spectropolarimetric observations identified twelve of this type of objects. These twelve objects constitute the complete sample of all known polar--scattered Seyfert~1 galaxies. X--ray studies of these distinctive {\\it polar--scattered} Seyfert~1 galaxies are a powerful tool to prove the basis of the model proposed by Smith et al. (2004) and therefore to further test the scheme of Unified Models for AGNs. We have analysed the eight objects included in the Smith et al. (2004) sample with available X-ray data. We present for the first time, X-ray analysis performed with \\xmm\\ of four them (Fairall 51, Mrk 704, ESO 323-G077, and IRAS 15091-2107). The results of the four remaining ones (Mrk 231, NGC 3227, Mrk 766, and NGC 4593) have been obtained from the literature. In the following section, we present the results on the analysis of the four objects observed by the first time by \\xmm. In Section~\\ref{sect:results}, we combine our results with the four already published {\\it polar--scattered} Seyfert 1 galaxies and address the conclusions of this work. ", "conclusions": "" }, "0710/0710.5330_arXiv.txt": { "abstract": "% {The observed abundance peculiarities of many chemical species relative to the expected cluster metallicity in blue horizontal-branch (BHB) stars presumably appear as a result of atomic diffusion in the photosphere. The slow rotation (typically $v\\sin{i}<$ 10 km s$^{-1}$) of BHB stars with effective temperatures $T_{\\rm eff}>$ 11,500~K supports this idea since the diffusion mechanism is only effective in a stable stellar atmosphere.} {In this work we search for observational evidence of vertical chemical stratification in the atmospheres of six hot BHB stars: B84, B267 and B279 in M15 and WF2-2541, WF4-3085 and WF4-3485 in M13.} {We undertake an abundance stratification analysis of the stellar atmospheres of the aforementioned stars, based on acquired Keck HIRES spectra.} {We have found from our numerical simulations that three stars (B267, B279 and WF2-2541) show clear signatures of the vertical stratification of iron whose abundance increases toward the lower atmosphere, while the other two stars (B84 and WF4-3485) do not. For WF4-3085 the iron stratification results are inconclusive. B267 also shows a signature of titanium stratification. Our estimates for radial velocity, $v\\sin{i}$ and overall iron, titanium and phosphorus abundances agree with previously published data for these stars after taking the measurement errors into account. The results support the hypothesis regarding the efficiency of atomic diffusion in the stellar atmospheres of BHB stars with $T_{\\rm eff}>$ 11,500~K. } {} ", "introduction": "According to the current understanding of stellar evolution, the horizontal-branch (HB) stars are post-main sequence stars that burn helium in their core and hydrogen in a shell (e.g. Moehler \\cite{Moehler04}). In this paper we consider the HB stars that are located in the blue part of the HB, to the left of the RR Lyrae instability strip. Most researchers call them blue horizontal-branch (BHB) stars to distinguish them from the red horizontal-branch (RHB) stars, which exhibit different observational properties. Sandage \\& Wallerstein (\\cite{S+W60}) have found from analysis of the colour-magnitude diagrams of globular clusters\\footnote{Most of the known BHB stars are found in globular clusters.} that the HB % generally becomes bluer with decreasing metallicity. Derived masses of the cool ($T_{\\rm eff}<$11,500~K) BHB stars in the globular cluster NGC~6388 (Moehler \\& Sweigart \\cite{Moehler+S06}) are in a good agreement with the predictions of canonical HB evolution, except for the hot BHB stars with $T_{\\rm eff}>$11,500~K, where the estimated stellar masses seem to be lower than the canonical values. \\begin{table*}[th] \\parbox[t]{\\textwidth}{ \\centering \\caption[]{Journal of Keck+HIRES spectroscopic observations of the selected hot BHB stars from Behr (\\cite{Behr03b}).} \\begin{tabular}{lcccccc} \\hline \\hline Cluster/Star& HJD &Exposure& S/N & Coverage & Seeing & Slit' width\\\\ &2450000+& Time (s) & & (\\AA) & (arcsec) & (arcsec) \\\\ \\hline M13/WF2-2541& 1046.7902 &3$\\times$1500& 44 & 3885-6292 & 0.90 & 0.86 \\\\ M13/WF4-3085& 1052.7338 &3$\\times$1200& 37 & 3888-5356 & 1.10 & 0.86\\\\ M13/WF4-3485& 1053.7793 &3$\\times$1200& 34 & 3888-5356 & 0.90 & 0.86\\\\ M15/B84 & 1053.9024 &4$\\times$1400& 34 & 3888-5356 & 0.80 & 0.86\\\\ M15/B267 & 1053.9731 &4$\\times$1400& 26 & 3888-5356 & 0.80 & 0.86\\\\ M15/B279 & 1053.8312 &4$\\times$1400& 34 & 3888-5356 & 0.90 & 0.86\\\\ \\hline \\end{tabular} } \\label{tab1} \\end{table*} The Hertzsprung-Russell diagrams of some globular clusters show long blue tails (an extension of the HB), populated by very hot BHB stars and extreme horizontal branch (EHB) stars. Published data on the BHB stars argue that the hot BHB stars show remarkable differences in physical properties when compared to the cool BHB stars. Using high-precision photometry of stars in M13, Ferraro et al. (\\cite{Ferraro+98}) have found gaps in the distribution of stars along the blue tail. One of these gaps, labeled as G1, is located at $T_{\\rm eff}\\sim$11,000-12,000~K. Grundahl et al. (\\cite{Grundahl+98}) used the results of Str\\\"{o}mgren $uvby\\beta$-photometry finding good agreement between the theoretical prediction of stellar evolution models and the observed location of BHB stars, except for the hot BHB stars, whose $u$-magnitudes are brighter than predicted. It appears that this $u$-jump is observed for the hot BHB stars and coincides with the temperature range of the G1 gap in M13. Similar $u$-jumps have also been found for other globular clusters (Grundahl et al. \\cite{Grundahl+99}). For the hot BHB with effective temperatures up to 20,000 K, the surface gravities derived from the fits of Balmer and helium line profiles appear to be lower than the predictions of stellar evolution models, while the gravities derived for the stars outside this temperature range are in good agreement with theoretical predictions (Moehler et al. \\cite{Moehler+95, Moehler+97a, Moehler+97b, Moehler+03}). The stellar rotation velocity distribution of BHB stars also appears to have a discontinuity at $T_{\\rm eff} \\simeq $ 11,500~K (Peterson et al.~\\cite{Peterson+95}; Behr et al. \\cite{Behr+00a}; Recio-Blanco et al.~\\cite{RB+04}), indicating that the hotter stars show modest rotation with $v\\sin{i}<$ 10 km s$^{-1}$, while the cooler stars rotate more rapidly on average. Comprehensive surveys of abundances also show that the hot BHB stars have abundance anomalies when compared to the cool BHB stars in the same cluster (Glaspey et al. \\cite{Glaspey+89}; Grundahl et al. \\cite{Grundahl+99}; Behr et al. \\cite{Behr+99, Behr+00b}, Behr \\cite{Behr03a}; Fabbian et al. \\cite{Fabbian+05}; Pace et al. \\cite{Pace+06}). The observed phenomena such as the low gravity, photometric jumps and gaps, abundance anomalies and slow rotation suggest that atomic diffusion could be important in the stellar atmospheres of hot BHB stars. Atomic diffusion arises from the competition between radiative acceleration and gravitational settling. This can produce a net acceleration on atoms and ions, which results in their diffusion through the atmosphere (Michaud~\\cite{Michaud70}). In order for atomic diffusion to produce a vertical stratification of the abundances of particular elements, the stellar atmosphere must be hydrodynamically stable. According to Landstreet (\\cite{Landstreet98}), photospheric convection should be very weak at the effective temperatures of BHB stars. Theoretical atmospheric models of Hui-Bon-Hoa, LeBlanc \\& Hauschildt (\\cite{Hui-Bon-Hoa+00}) showed that the observed photometric jumps and gaps for hot BHB stars can be explained by elemental diffusion in their atmosphere. Behr (\\cite{Behr03b}) has shown that adoption of a microturbulent velocity of 0 or 1 km s$^{-1}$ provides the best fit to line strengths in the spectra of hot BHB stars. This fact supports the proposal that strong velocity fields are not present in the atmospheres of hot BHB stars. While synthesizing spectral line profiles, Khalack et al. (\\cite{Khalack+07}) have recently found vertical abundance stratification of sulfur in the atmosphere of the field BHB star HD~135485. In this paper we also attempt to detect signatures of vertical abundance stratification of elements from line profile analyses of several other BHB stars for which we have appropriate data. Together with the data on stratification of the sulfur abundance in HD~135485, new positive results would provide a convincing argument in favour of efficient atomic diffusion in the atmospheres of hot BHB stars. In Sec.~\\ref{obs} we discuss the properties of the acquired spectra, while in Sec.~\\ref{mod} we describe details concerning the simulation routine and adopted atmospheric parameters for the program stars. The evidence for vertical stratification of some chemical species is given in Sec.~\\ref{vert}, while the estimation of mean abundances and velocities is described in Sec.~\\ref{mean}. A discussion follows in Sec.~\\ref{discuss}. ", "conclusions": "\\label{discuss} In this paper we continue our attempts to detect vertical abundance stratification in the atmospheres of BHB stars. After the report by Bonifacio et al. (\\cite{Bonifacio+95}) of the vertical stratification of helium in the atmosphere of Feige 86, it became clear that the abundances of other chemical species may also be stratified. We devoted special interest to iron because hot BHB stars usually have an enhanced iron abundance (e.g. Behr~\\cite{Behr03b}), suggesting that this element may be strongly affected by diffusion. Analysing the spectra of another hot BHB star HD~135485 (Khalack et al. \\cite{Khalack+07}) we did not find direct evidence of iron stratification, but revealed strong signatures of sulfur depletion in the deeper atmospheric layers. However, HD~135485 is different from the other BHB stars in that its spectrum shows evidence of helium enrichment (in comparison with the solar abundance), while in the atmospheres of the other BHB stars helium is depleted. Therefore, we directed our attention to BHB stars where the iron abundance is near the solar abundance or enhanced, and helium is depleted. The results obtained argue that at least three stars (B267 and B279 in M15 and % WF2-2541 in M13) show clear signatures of vertical stratification of their iron abundance, while for WF4-3085 the results are suggestive, but not conclusive. The other two stars studied here (B84 in M15 and WF4-3485 in M13) do not show stratification of iron and their averaged iron abundance is close to solar (but is enhanced in comparison with its cluster value). B267 shows also a signature of vertical stratification of titanium (see Fig.~\\ref{B267}b). Since our simulations show that the turnup feature observed in the iron stratification profile is strongly dependent on microturbulent velocity, the value of the abundance at low optical depths is uncertain. Of course, if the theoretical framework supposes that these abundance gradients are due to atomic diffusion, microturbulence should be weak since a stable atmosphere is needed for diffusion to be dominant. It should be noted that for corresponding optical depths the abundance profile of iron is similar in the three BHB stars that exhibit stratification. The reason that the other two stars in our study do not show clear signs of iron stratification (B84 and WF4-3485) might be related to evolutionary effects or the presence of other competing hydrodynamical processes. The absence of Ti\\,{\\sc ii} and P\\,{\\sc ii} lines in their spectra might be evidence of this. In conclusion, the results shown here add to the mounting evidence of the existence of vertical abundance stratification, and hence atomic diffusion, in the atmospheres of BHB stars." }, "0710/0710.5106_arXiv.txt": { "abstract": "{ The excess of diffuse galactic gamma rays above 1 GeV, as observed by the EGRET telescope on the NASA Compton Gamma Ray Observatory, shows all the key features from Dark Matter (DM) annihilation: (i) the energy spectrum of the excess is the same in all sky directions and is consistent with the gamma rays expected for the annihilation of WIMPs with a mass between 50-100 GeV; (ii) the intensity distribution of the excess in the sky is used to determine the halo profile, which was found to correspond to the usual profile from N-body simulations with additional substructure in the form of two doughnut-shaped structures at radii of 4 and 13 kpc; (iii) recent N-body simulations of the tidal disruption of the Canis Major dwarf galaxy show that it is a perfect progenitor of the ringlike Monoceros tidal stream of stars at 13 kpc with ring parameters in agreement with the EGRET data; (iiii) the mass of the outer ring is so large, that its gravitational effects influence both the gas flaring and the rotation curve of the Milky Way. Both effects are clearly observed in agreement with the DMA interpretation of the EGRET excess. } % ", "introduction": "If dark matter (DM) is created thermally during the Big Bang the present relic density is inversely proportional to $\\langle\\sigma v\\rangle$, the annihilation cross section $\\sigma$ of DM particles, usually called WIMPS (Weakly Interacting Massive Particles), times their relative velocity. The average is taken over these velocities. This inverse proportionality is obvious, if one considers that a higher annihilation rate, given by $\\langle\\sigma v\\rangle n_\\chi$, would have reduced the relic density before freeze-out, i.e. the time, when the expansion rate of the Universe, given by the Hubble constant, became equal to or larger than the annihilation rate. For the present value of $\\Omega h^2=0.105 \\pm 0.008$, as measured by WMAP \\cite{wmap}, the thermally averaged total cross section at the freeze-out temperature of $m_\\chi/22$ must have been around $3\\cdot 10^{-26} ~{\\rm cm^3s^{-1}}$ \\cite{jungman}. If the s-wave annihilation is dominant, as expected in many supersymmetric models, then the annihilation cross section is energy independent, i.e. the cross section given above is also valid for the cold temperatures of the present universe \\cite{susy}. Such a large cross section will lead to a production rate of mono-energetic quarks\\footnote{The quarks are mono-energetic, since the kinetic energy of the cold dark matter particles is expected to be negligible with the mass of the particles, so the energy of the quarks equals the mass of the WIMP.} in our Galaxy, which is 40 orders of magnitude above the rate produced at any accelerator. The fragmentation of these quarks will lead to a large flux of gamma rays with a characteristic energy spectrum quite different from the background of cosmic ray interactions with the interstellar material of the Galaxy. In addition, gamma rays have the advantage that they point back to the source and do not suffer energy losses, so they are the ideal candidates to trace the dark matter density. The charged components interact with Galactic matter and are deflected by the Galactic magnetic fields, so they do not point back to the source. Therefore the charged particle fluxes have large uncertainties from the pro\\-pagation models, which determine how many of the produced particles arrive at the detector. For gamma rays the propagation is straightforward: only the ones pointing towards the detector will be observed. An excess of diffuse gamma rays compatible with dark matter annihilation (DMA) has indeed been observed by the EGRET telescope on board of NASA's CGRO (Compton Gamma Ray Observatory)\\cite{us}. The excess was observed in all sky directions, which would imply that DM is not dark anymore, but shining in gamma rays. Of course, such an important observation needs to be scrutinized heavily. Before discussing the criticism the evidence and new confirmation from N-body simulations and the gas flaring is presented in the next section. \\begin{figure} \\begin{center} \\vspace*{-0.65cm} \\includegraphics [width=0.42\\textwidth,clip]{Plots/fig2.eps} \\caption[]{Fit of the shapes of background and DMA signal to the EGRET data in the direction of the Galactic centre The light shaded (yellow) area indicates the background using the shapes known from accelerator experiments, while the dark shaded (red) area corresponds to the signal contribution from DMA for a 60 GeV WIMP mass. The intermediate (blue) shaded area corresponds to a variation of the WIMP mass between 50 and 70 GeV.} \\label{fig2} \\end{center} \\end{figure} ", "conclusions": "" }, "0710/0710.1808_arXiv.txt": { "abstract": "Observations of a large solar flare of December 13, 2006, using Solar Optical Telescope (SOT) on Hinode spacecraft revealed high-frequency oscillations excited by the flare in the sunspot chromosphere. These oscillations are observed in the region of strong magnetic field of the sunspot umbra, and may provide a new diagnostic tool for probing the structure of sunspots and understanding physical processes in solar flares. ", "introduction": "Solar flares represent a process of conversion of magnetic energy into heat, kinetic energy of plasma eruptions and high-energy particles. Solar flares may excite various types of oscillations and waves in various layers of the Sun, from \"sunquakes\" in the interior \\citep{Kosovichev1998} to coronal Morton waves \\citep{Moreton1960} and coronal loop oscillations \\citep{Aschwanden1999}. The mechanisms of these oscillations and waves are not fully understood yet, but obviously related to the energy release and transport properties. For instance, the seismic response (\"sunquakes\") are believed to be related to the hydrodynamic impact on the lower atmosphere and photosphere by shocks generated in the area heated by high-energy electrons. The spatial-temporal properties of the seismic wave source are closely related to the properties of hard X-ray source \\citep{Kosovichev2006a,Kosovichev2006b}. The sunquakes represent packets of high-frequency acoustic waves traveling through the Sun's interior. The waves propagate through the active regions and sunspots with strong magnetic field without significant distortion of the wave front and large changes in the travel times. The magnetic field effects in the sunquake waves, which to some extent are obviously present, have not been detected. Here, we report on observations from Hinode spacecraft of a different type of oscillations excited by a solar flare. These oscillations are observed in the chromosphere of the sunspot umbra and inner penumbra, and, thus, probably represent some kind of magnetohydrodynamic oscillatory modes. Unfortunately, the relatively low cadence of the observations did not allow us to identify the specific mode of the oscillations. Nevertheless, these oscillations carry potentially interesting information about the flare energy release and transport and properties of sunspots, and deserve further observational and theoretical studies. ", "conclusions": "The observations of the solar flare of December 13, 2006, from Hinode reveal a new type of flare-excited oscillations. The oscillations observed in Ca~II~H images appeared in the chromospheric layers of the sunspot umbra immediately after the impulsive phase. They had the amplitude 2--4 times larger than the pre-flare oscillations in the umbra. Also, their frequency seemed to be higher. There is a weak evidence that during the first 30-40 min the oscillations represent waves traveling through the umbra in the direction away from the flare ribbon with a speed of 50--100 km/s. Then, the oscillation become more irregular with some occasional wave packets. The lifetime of these oscillations is probable about 8 hours. The uncertainties in the data analysis and interpretation are caused by the low cadence of the Hinode observing program during the flare. The images were taken every 2 min while the characteristic oscillation period is about 3 min. Most of the oscillation power is probably even at the shorter periods. Thus, for detailed studies of these oscillations it will be important to increase the image cadence. The image cadence should be sufficiently high to capture the initial waves traveling with a speed of $\\sim 100$ km/s according to our preliminary estimates. This speed indicates that the waves are of an MHD type, and if their speed is of the order of magnitude of the Alfven speed then they should propagate rather low in the sunspot chromosphere. Thus, simultaneous observations in photospheric lines would be interesting. The oscillation amplitude was several times higher than the amplitude of preflare umbral oscillations, which can be as high as 5--6 km/s \\citep{Yoon1995}, and thus may reach supersonic velocities of 10--20 km/s. Sunspot oscillations have been studied intensively for many years (for a review see \\citet{Staude1999}) but these Hinode observations seem to be first that show enhanced oscillations in the umbra, associated with a solar flare. Further investigations of these oscillations are of great interest for understanding the processes in solar flares and sunspots." }, "0710/0710.1039_arXiv.txt": { "abstract": "{Turbulent motions in stellar convection zones generate acoustic energy, part of which is then supplied to normal modes of the star. Their amplitudes result from a balance between the efficiencies of excitation and damping processes in the convection zones. } {We develop a formalism that provides the excitation rates of non-radial global modes excited by turbulent convection. As a first application, we estimate the impact of non-radial effects on excitation rates and amplitudes of high-angular-degree modes which are observed on the Sun. } {A model of stochastic excitation by turbulent convection has been developed to compute the excitation rates, and it has been successfully applied to solar radial modes (Samadi \\& Goupil 2001, Belkacem et al. 2006b). We generalize this approach to the case of non-radial global modes. This enables us to estimate the energy supplied to high-($\\ell$) acoustic modes. Qualitative arguments as well as numerical calculations are used to illustrate the results. } {We find that non-radial effects for $p$~modes are non-negligible: \\\\ - for high-$n$ modes (i.e. typically $n > 3$) and for high values of $\\ell$; the power supplied to the oscillations depends on the mode inertia.\\\\ - for low-$n$~modes, independent of the value of $\\ell$, the excitation is dominated by the non-diagonal components of the Reynolds stress term. } {We carried out a numerical investigation of high-$\\ell$ $p$~modes and we find that the validity of the present formalism is limited to $\\ell < 500$ due to the spatial separation of scale assumption. Thus, a model for very high-$\\ell$ $p$-mode excitation rates calls for further theoretical developments, however the formalism is valid for solar $g$~modes, which will be investigated in a paper in preparation. } ", "introduction": "\\label{intro} Amplitudes of solar-like oscillations result from a balance between stochastic excitation and damping in the outermost layers of the convection zone, which extends to near the surface of the star. Accurate measurements of the rate at which acoustic energy is supplied to the solar $p$~modes are available from ground-based observations (GONG, BiSON) as well as from spacecraft (SOHO/GOLF and MDI). From those measurements and a comparison with theoretical models, it has been possible to demonstrate that excitation is due to eddy motions in the uppermost part of the convection zone and by advection of entropy fluctuations. Stochastic excitation of \\emph{radial} modes by turbulent convection has been investigated by means of several semi-analytical approaches \\citep{GK77,GK94,B92,Samadi00I}; they differ from each other in the nature of the assumed excitation sources, the assumed simplifications and approximations, and also by the way the turbulent convection is described \\citep[see reviews by][]{Stein04,Houdek06}. Two major mechanisms have nevertheless been identified to drive the resonant $p$~modes of the stellar cavity: the first is related to the Reynolds stress tensor and as such represents a mechanical source of excitation; the second is caused by the advection of turbulent fluctuations of entropy by turbulent motions (the so-called ``entropy source term'') and as it such represents a thermal source of excitation \\citep{GK94,Stein01B}. Samadi \\& Goupil (2001, hereafter Paper~I) proposed a generalized formalism, taking the Reynolds and entropy fluctuation source terms into account. In this model, the source terms are written as functions of the turbulent kinetic energy spectrum and the temporal-correlation function. This allows us to investigate several possible models of turbulence \\citep{Samadi02II,Samadi02I}. The results have been compared with GOLF data for radial modes, and the theoretical values are found to be in good agreement with the observations \\citep{Samadi02I}. Part of the remaining discrepancies has been recently removed by taking into account the asymmetry introduced by turbulent plumes \\citep{Belkacem06a,Belkacem06b}. In this paper we take an additional step in extending the \\cite{Samadi00I} formalism to the case of non-radial global modes. This will enable us to estimate the excitation rates for a wide variety of $p$ and $g$ modes excited in different types of stars. The present model provides the energy supplied to the modes by turbulence in inner, as well as outer, stellar convective regions, provided the turbulent model appropriate for the relevant region is used. Studies of stochastic excitation of solar radial modes \\citep{Samadi02II,Samadi02I} have given us access mainly to the radial properties of turbulence. The present generalized formalism enables us to take into account the horizontal properties of turbulence (through the non-radial components of the Reynolds stress) in the outermost part of the convective zone. In the Sun, high-angular-degree $p$~modes (as high as one thousand) have been detected \\citep[e.g.,][]{Rabello04}. From an observational point of view, \\cite{Woodard01} found that the energy supplied to the mode increases with $\\ell$, but that above some high-$\\ell$ value, which depend on the radial order $n$ \\citep[see][Fig.~2]{Woodard01}, the energy decreases with increasing $\\ell$. They mentioned the possibility of an unmodelled mechanism of damping. Hence one of the motivations of this work is to investigate such an issue. As a first step, we develop here a theoretical model of the stochastic excitation taking into account the $\\ell$-dependence of the source terms to seek a physical meaning for such a behaviour of the amplitudes. Modelling of the mechanisms responsible for the excitation of non-radial modes is not only useful for high-$\\ell$ acoustic modes but also for gravity modes, which are intrinsically non-radial. As for $p$ modes, $g$ modes are stochastically excited by turbulent convection; the main difference is that the dominant restoring force for $g$ modes is buoyancy. We however stress that convective penetration is another possible excitation mechanism for $g$~modes \\citep[e.g.][]{Dintrans05}. Such modes are trapped in the radiative interior of the Sun, so their detection promises to give a much better knowledge of the deep solar interior. However, such modes are evanescent in the convection zone; their amplitudes at the surface are very small and their detection remains controversial. A theoretical prediction of their amplitudes is thus an important issue. It requires an estimation of the excitation rates but also of the damping rates. Unlike $p$~modes, the damping rates cannot be inferred from observations and this introduces considerable uncertainties; e.g. theoretical estimates of the $g$-mode amplitudes \\citep{Gough85,Kumar96} differ from each other by orders of magnitudes, as pointed out by \\cite{CD2002}. We thus stress that the present work focuses on the excitation rates -- damping rates are not investigated. A specific study of gravity modes will be considered in a forthcoming paper. The paper is organized as follows: Sect.~2 introduces the general formalism, and a detailed derivation of the Reynolds and entropy source terms is provided. In Sect.~3, we demonstrate that the formalism of \\cite{Samadi00I} is a special case and an asymptotic limit of the present model. In Sect.~4, we determine, using qualitative arguments, the different contributions to the excitation rates and identify the dominant terms involving the angular degree (\\,$\\ell$\\,). Sect.~5 presents the numerical results: excitation rates are presented and discussed. Sect.~6 discusses the limitations of the model and some conclusions are formulated in Sect.~7. ", "conclusions": "\\label{discussion} \\subsection{The separation of scales} \\label{sep_ech} The main assumption in this general formalism appears in \\eq{mean_square_amplitude_trapping}, where it has been assumed that the spatial variation of the eigenfunctions is large compared to the typical length scale of turbulence, leading to what we call \\emph{the separation of scales}. In order to test this assumption, one must compare the oscillation wavelength to the turbulent one or equivalently the wavenumbers. To this end, we use the dispersion relation \\cite[see][]{Unno89} \\begin{equation} \\label{dispersion} k_r^2 = \\frac{\\omega^2}{c_s^2}(1 - \\frac{S_\\ell^2}{\\omega^2})(1 - \\frac{N^2}{\\omega^2}) \\quad \\mbox{ and} \\quad k_h^2 = \\frac{L^2}{r^2} \\end {equation} where $N$ is the buoyancy frequency, $S_\\ell$ the Lamb frequency and $k_r,k_h$ the radial and horizontal oscillation wavenumbers, respectively, and $L^2 = \\ell(\\ell+1)$.\\\\ For the turbulent wavenumber, we choose to use, as a lower limit, the convective wavenumber $k_{conv} = 2 \\pi / L_c$ where $L_c$ is the typical convective length scale. Thus, the assumption of separation of scales is fulfilled provided \\begin{equation} k_{r,h} / k_{conv} \\ll 1 \\end{equation} In Figs.~\\ref{sep_radial} are plotted the ratios $k_r/k_{conv}$ and $k_h/k_{conv}$ respectively. Those plots focus on the uppermost part of the solar convection zone where most of the excitation takes place. The assumption of separation of scale is valid for the horizontal component of the oscillation since one has $k_h / k_{conv} \\ll 1$ (for $\\ell \\leq 500$) in the region where excitation is dominant. However, we must recall that our criterion is based on the mixing length to compute $k_{conv}$. As shown by \\cite{Samadi02II} using 3D numerical simulations, the convective length scale (computed using the CESAM code, see Sect.~\\ref{calculation_method}) must be multiplied by a factor around five in order to reproduce the injection scale ($L_c$) in the superadiabatic layers. Hence, for a more conservative criterion, we must then multiply the ratio $k_h / k_{conv}$ by a factor of five, this leads to a ratio near unity for $\\ell \\approx 500$ (see Fig.~\\ref{sep_radial}). Thus, for higher values of the angular degree, the separation-of-scale hypothesis becomes doubtful. \\begin{figure}[t] \\begin{center} \\includegraphics[height=6cm,width=9cm]{sep_horizontal.eps}\\\\ \\includegraphics[height=6cm,width=9cm]{sep_radial.eps} \\caption{{\\bf Top:} Ratio of the horizontal oscillation wavenumber to the convective wavenumber ($k_h / k_{conv}$), versus the normalized radius ($r/R$). $k_{conv}$ is computed using the mixing length theory such that $k_{conv} = 2 \\pi / L_c$ ($L_c$ is the mixing length) and $k_{r}$ is computed using the dispersion relation \\eq{dispersion}. Note that the ratio $k_h / k_{conv}$ is computed for a frequency of around $\\nu= 3$ mHz, depending of the angular degree (\\,$\\ell$\\,). {\\bf Bottom:} the same as in the top but for the ratio $k_r / k_{conv}$.} \\label{sep_radial} \\end{center} \\end{figure} Concerning the radial component of the oscillation wavenumber, the limiting value of $\\ell$ seems to be the same (i.e. $\\ell=500$). Thus, we conclude that for modes of angular degree lower than 500 one can use the separation of scales assumption. For $\\ell >$~500 the characteristic length of the mode becomes smaller than the characteristic length $L_c$ of the energy bearing eddies. Those modes will then be excited by turbulent eddies with length-scale smaller than $L_c$, i.e. lying in the turbulent cascade. These eddies inject less energy into the mode than the energy bearing eddies do, since they have less kinetic energy. We can then expect that -- at fixed frequency -- they received less energy from the turbulent eddies than the low-degree modes. A theoretical development is currently underway to properly treat the case of very high $\\ell$ modes. \\subsection{The closure model} A second approximation in the present formalism is the use of a closure model. The uppermost part of the convection zone is a turbulent convective system composed of two flows (upward and downward) and the probability distribution function of the fluctuations of the vertical velocity and temperature does not obey a Gaussian law \\citep{Lesieur97}. Thus, the use of the quasi-normal approximation (QNA, \\citet{Million41}) which is exact for a normal distribution, is no longer rigorously correct. A more realistic closure model has been developed in \\cite{Belkacem06a} and can be easily be adapted for high-$\\ell$ modes. This alternative approach takes into account the existence of two flows (the up- and downdrafts) within the convection zone. However, the QNA is nevertheless often used for the sake of simplicity as is the case here. Note that, when using the closure model with plumes, it is no longer consistent to assume that the third-order velocity moments strictly vanish, however as shown by \\cite{Belkacem06a,Belkacem06b}, their contribution is negligible in the sense that their effect is smaller than the accuracy of the presently available observational data. \\subsection{Mode inertia} We have shown that the excitation rates for high-$\\ell$ and $n$ modes are sensitive to the variation of the mode inertia ($I$). $I$ depends on the structure of the stellar model and the properties of the eigenfunctions in these external regions. \\cite{Samadi06} have shown that different local formulations of convection can change the mode inertia by a small amount. This sensitivity then affects the computed excitation rates ($P$). However, the changes induced in $P$ are found to be smaller than the accuracy to which the mode excitation rates are derived from the current observations \\citep[see][]{Baudin05,Belkacem06b}. Furthermore, concerning the way the modes are obtained, we have computed non-adiabatic eigenfunctions using the time-dependent formalism of Gabriel for convection \\citep[see][]{grigahcene05}. The mode inertia obtained with these non-adiabatic eigenfunctions exhibits a $\\nu$ dependency different from those obtained using adiabatic eigenfunctions (the approximation adopted in the present paper). On the other hand, the mode inertia using non-adiabatic eigenfunctions \\citep[see][for details]{Houdek1999} obtained according to Gough's time-dependent formalism of convection \\citep{Gough77} shows smaller differences with the adiabatic mode inertia. Accordingly, the way the interaction of oscillation and time-dependent convection is modelled affects the eigenfunctions differently. As explained in Sect.~\\ref{surf}, the formalism developed in this paper can be an efficient tool to derive constraints on the mode inertia to discriminate between the different treatments of convection. Further work is thus needed on that issue." }, "0710/0710.3499_arXiv.txt": { "abstract": "High-resolution ultraviolet observations of the black hole X-ray binary Cygnus X-1 were obtained using the Space Telescope Imaging Spectrograph on the Hubble Space Telescope. Observations were taken at two epochs roughly one year apart; orbital phase ranges around $\\phi_{orb}$ = 0 and 0.5 were covered at each epoch. We detect P Cygni line features from high (N~V, C~IV, Si~IV) and absorption lines from low (Si~II, C~II) ionization state material. We analyze the characteristics of a selection of P Cygni profiles and note, in particular, a strong dependence on orbital phase for the high ionization material: the profiles show strong, broad absorption components when the X-ray source is behind the companion star and noticeably weaker absorption when the X-ray source is between us and the companion star. We fit the P~Cygni profiles using the Sobolev with Exact Integration method applied to a spherically symmetric stellar wind subject to X-ray photoionization from the black hole. Of the wind-formed lines, the Si\\,IV doublet provides the most reliable estimates of the parameters of the wind and X-ray illumination. The velocity $v$ increases with radius $r$ (normalized to the stellar radius) according to $v=v_\\infty(1-r_\\star/r)^\\beta$, with $\\beta\\approx0.75$ and $v_\\infty\\approx1420$~km~s$^{-1}$. The microturbulent velocity was $\\approx160$~km~s$^{-1}$. Our fit implies a ratio of X-ray luminosity (in units of 10$^{38}$ erg s$^{-1}$) to wind mass-loss rate (in units of 10$^{-6} \\msun~yr^{-1}$) of L$_{X,38}/\\dot M_{-6} \\approx 0.33$, measured at $\\dot M_{-6}$ = 4.8. The lines from the lower ionization species and the He~II~$\\lambda$1640 absorption are consistent with formation in the photosphere of the normal companion. Our models determine parameters that may be used to estimate the accretion rate onto the black hole and independently predict the X-ray luminosity. Our predicted L$_x$ matches that determined by contemporaneous RXTE ASM remarkably well, but is a factor of 3 lower than the rate according to Bondi-Hoyle-Littleton spherical wind accretion. We suggest that some of the energy of accretion may go into powering a jet. ", "introduction": "Cygnus X-1, the only Galactic X-ray binary with a high mass companion where existing observations require a black hole for the compact object, was first discovered in 1962 (Cowley 1992, and references therein). In addition to the well-established high mass function found with optical observations, the X-ray data of Cyg X-1 display transitions from a high flux state in the 2-10 keV band (where a strong soft component dominates) to a low flux state (where the soft component largely disappears) that has been interpreted as characteristic of black hole systems. It also displays highly broadened Fe K$\\alpha$ emission (Miller~\\etal~2002) that is consistent with models for X-ray reflection in Galactic black holes and AGNs. The broad-line shape of Fe K$\\alpha$ may be caused by Doppler shifts and the gravitational field of the black hole. The Cyg X-1 system consists of a supergiant star and a compact object. The mass of the compact object is in the range 7-20 $\\msun$ (Shaposhnikov \\& Titarchuk 2007; Ziolkowski 2005) the mass of the visible star is in the range 18-40$\\msun$ (Ziolkowski 2005; Tarasov~\\etal~2003; Brocksopp~\\etal~1999). The binary orbital period is 5.6 days (Bolton 1972). Miller~\\etal~(2005) using Chandra/HETG observations find that the X-ray spectrum of Cygnus X-1 at phase 0.76 is dominated by absorption lines, in strong contrast to spectra of other HMXBs such as Vela X-1 and Cen X-3. Schulz~\\etal~(2002) report marginal evidence for ionized Fe transitions with P-Cygni type profiles at orbital phase 0.93 whereas Marshall~\\etal~(2001) find no evidence for such line profiles at phase 0.84. Miller~\\etal~(2005) suggest that, while the spectra of Cen X-3 can be modelled by a spherically-symmetric wind (Wojdowski~\\etal~2003), the X-ray absorption spectrum of Cyg X-1 requires dense material preferentially along the line of sight; considered together, the Chandra spectra provide evidence in X-rays for a focused wind in Cygnus X-1. Our Space Telescope Imaging Spectrometer (STIS) observations of Cygnus X-1 were obtained when Cyg X-1 was in its soft/high X-ray state and show line profiles that change significantly between orbital phases 0.0 and 0.5. We interpret these changes in terms of models that include the effects of X-ray photoionization on the stellar wind of the normal companion or of a focused wind. We test our model predictions with contemporaneous X-ray observations. The observations and analysis are described in Section 2, our models of the line profiles are presented in Section 3, and our interpretation and conclusions are discussed in Section 4. \\\\ ", "conclusions": "The Space Telescope Imaging Spectrograph on Hubble provides the highest resolution ultraviolet spectra taken of Cyg X-1 to date. Observations were taken at two epochs roughly a year apart: at each epoch orbital phases when the compact object is behind the stellar companion and when the compact object is in front of the companion star were covered. We find P~Cygni profiles from high ionization (N\\,V, C\\,IV, Si\\,IV) gas. For both epochs the P~Cygni profiles show significantly less absorption at phases when the compact object is in the line of sight. RXTE observations indicate that the X-ray flux of the system was at a similar level at each epoch. The observed changes can be attributed entirely to orbital effects. We interpret this to mean that X-rays from the compact object photoionize the wind from the massive companion resulting in reduced absorption by the wind material. P Cygni profiles of selected species are consistent with the Hatchett-McCray effect, in which X-rays from the compact object photoionize the stellar wind from the companion star, thereby reducing absorption. This effect also appears in UV observations of LMC X-4 and SMC X-1 (Boroson et al. 1999; Vrtilek et al. 1997; Treves et al. 1980). SEI models can fit the observed P Cygni profiles and provide measurements of the stellar wind parameters. The Si\\,IV fits are the most reliable and we use them to determine L$_x/\\dot{M} =4.8 \\pm0.3 \\times 10^{42}$ ergs s$^{-1}$ M$_{\\odot}^{-1}$ yr, where L$_{X}$ is the X-ray luminosity and $\\dot{M}$ is the mass-loss rate of the star. The results from the C\\,IV and N\\,V lines are less reliable because they are saturated and the C\\,IV fit does not match the data well. For these fits we fixed the terminal velocity $\\nu_{\\infty}$ and the microturbulent velocity to those given by our fits to Si\\,IV. The best fit values for the optical depth in the ambient wind are high ($\\ge 10$). Once saturated, the OB star wind lines hardly change with large optical depth, but when much of the wind is ionized by the black hole, the ion fraction in the remaining regions can have a significant effect on the line profile. The reduced absorption when the compact object is in the line of sight is inconsistent with focusing of the wind toward the compact object, as has been suggested by several authors (e.g, Sowers~\\etal~1998; Tarasov~\\etal~2003; Miller~\\etal~2005), as then we would expect more absorbers in the line-of-sight and hence increased P Cygni absorption. Also, IUE observations taken at 8 orbital phases show a continuous variation in the P Cygni profiles with maximum absorption at phase 0.0 and minimum at 0.5 (Treves~\\etal~1980; van Loon~\\etal~2001). Further high spectral resolution ultraviolet observations of Cyg X-1 will be necessary to study the behavior of the P Cygni lines during different X-ray states. In an analysis of 2 years of RXTE/ASM data Wen~\\etal~(1999) found the 5.6 day orbital period of Cyg X-1 during the X-ray low/hard state, but no evidence of the orbital period during the high/soft state. Wen \\etal~suggest that absorption of X-rays by a stellar wind from the companion star can reproduce the observed X-ray orbital modulations in the hard state: The lack of modulation in the soft state could be due either to a reduction of the wind density during the soft state or to partial covering of a central hard X-ray emitting region by an accretion stream. Gies~\\etal~(2003) used the results of a four year spectroscopic monitoring program of the H$\\alpha$ emission strength of HDE226868, the normal companion to Cyg X-1, to argue that the low/hard X-ray state occurs when there is a strong, fast wind and accretion rate is low, while in the high/soft state a weaker, highly ionized wind attains only a moderate velocity and the accretion rate increases. The interpretations of both Wen~\\etal~(1999) and Gies~\\etal~(2003) are inconsistent with the fact that the {\\it total} X-ray luminosity from 1.3-200 keV remains constant during both the X-ray soft and hard states (Wen~\\etal~1999): the designation of X-ray high or X-ray low during these states is only applicable for the 1.5-12 keV ASM band. Since the 1.3-200 keV X-ray luminosity is unchanged from the hard to soft state, fluctuations in the narrow ASM band cannot be due to reduction in accretion, or obscuration of the X-ray source; rather it is a physical change that causes the dominant emission mechnism to switch from thermal to power-law. We note that the orbital modulation observed by Wen~\\etal~in the hard state ($\\pm 1.6$ ASM cts/sec around the average) is less than the errors on the counts during the soft state ($\\pm 3$ ASM cts/sec); we suggest that the quality of the RXTE ASM is not sufficient to detect this low level orbital modulation during the soft state. Ultraviolet observations clearly show orbital modulation during both X-ray hard and soft states (Gies~\\etal~2007). Our wind models explain both the X-ray and ultraviolet flux during the soft/high state. We need high spectral resolution ultraviolet observations during the hard/low state to determine if there is a change in wind density between states. Our determination of the mass-accretion rate can be considered as a positive check on the Hatchett-McCray models, as the Si\\,IV line gave $L_x=1.6\\times10^{37}$~erg~s$^{-1}$, and the model is subject to systematic uncertainties. However, this value of $L_x$ is a factor of 3 lower than the best estimate of the accretion rate according to Bondi-Hoyle-Littleton spherical wind accretion. We suggest that some of the energy of accretion may go into powering the jet. Our test of the dependence of our results on X-ray luminosity confirms the utility of our models and demonstrates that the wind outside the shadow zone is still sensitive to X-ray illumination, an arrangement which allows us to fit $L_{\\rm x}$ as a free parameter in our model. In future observations of the time-variability of the wind lines, the light travel-time effects may be used to advantage, as the wind may act as a ``low-pass filter\" to the X-ray observations, with the filter cutoff indicating the size of the ionized region (Kallman, McCray, \\&\\ Voit 1987)." }, "0710/0710.3450_arXiv.txt": { "abstract": "We performed a timing analysis of the 2003 outburst of the accreting X-ray millisecond pulsar \\xtejb\\ observed by RXTE. Using recently refined orbital parameters we report for the first time a precise estimate of the spin frequency and of the spin frequency derivative. The phase delays of the pulse profile show a strong erratic behavior superposed to what appears as a global spin-up trend. The erratic behavior of the pulse phases is strongly related to rapid variations of the light curve, making it very difficult to fit these phase delays with a simple law. As in previous cases, we have therefore analyzed separately the phase delays of the first harmonic and of the second harmonic of the spin frequency, finding that the phases of the second harmonic are far less affected by the erratic behavior. In the hypothesis that the second harmonic pulse phase delays are a good tracer of the spin frequency evolution we give for the first time a estimation of the spin frequency derivative in this source. The source shows a clear spin-up of $\\dot \\nu = 2.5(7) \\times 10^{-14}$ Hz sec$^{-1}$ (1 $\\sigma$ confidence level). The largest source of uncertainty in the value of the spin-up rate is given by the uncertainties on the source position in the sky. We discuss this systematics on the spin frequency and its derivative. ", "introduction": "Binary systems in which one of the two stars is a neutron star (NS hereafter) are among the most powerful X-ray sources of our Galaxy. The emission of X-rays is due to the matter transferred from the companion star and accreted onto the NS, and to the release of the immense gravitational energy during the fall or in the impact with the NS surface. A sub-category of such systems is called Low Mass X-ray Binaries (LMXB). LMXBs are characterized by low NS superficial magnetic fields ($< 10^9~ \\mathrm{Gauss}$) and by the low-mass ($< 1 \\, M_\\odot$) of the companion star. The so-called recycling scenario \\citep[see for a review][]{Bhatta_91} sees in the millisecond radio pulsars the last evolutionary step of LMXBs, where the torques due to the accretion of matter and angular momentum, together with the relatively weak magnetic fields, are able to spin-up the NS to periods of the order of one millisecond. When the accretion phase terminates and the companion star stops transferring matter, the NS can switch on as a millisecond radio pulsar, although no example has been reported yet. The recycling scenario received the long awaited confirmation only in 1998 with the discovery of the first millisecond X-ray pulsar in a transient LMXB; the first LMXB observed to show coherent pulsations at a frequency of $\\sim 400$ Hz was \\saxj\\ \\citep{Wijnands_98}, in which the NS is orbiting its companion star with a period of $\\sim 2.5$ hr \\citep{Chakra_98}. Why millisecond X-ray pulsars were so elusive is an argument still debated in literature. A possible reason can be due to the relatively low magnetic fields of these sources which has therefore less capability to channel the accreting matter onto the polar caps, then the chance to see a pulsed emission from a LMXB is quite low \\citep[see e.g.][]{Vaughan_94}, especially at high accretion rates. However, to date 10 LMXBs have been discovered to be accreting millisecond pulsars (see \\citealp{Wijnands_05}~for a review on the first 6 discovered, for the last four see \\citealp{Kaaret_06, Krimm_07, Casella_07, Altamirano_07}), and all of them are in transient systems. They spend most of the time in a quiescent state, with very low luminosities (of the order of $10^{31} - 10^{32}$ ergs/s) and rarely they go into an X-ray outburst with luminosities in the range $10^{36} - 10^{37}$ ergs/s. Although the recycling scenario seems to be confirmed by these discoveries, from timing analysis of accreting millisecond pulsars we now know that some of these sources show spin-down while accreting~ \\citep{Galloway_02, Papitto_07}. This means that it is of fundamental importance to study the far from being understood details of the mechanisms regulating the exchange of angular momentum between the NS and the accreting matter, and chiefly the role of the magnetic field in this exchange. The main way to do this is the study of the pulse phase shifts and their relations with other physical observable parameters of the NS. The pulse phase shifts are frequently affected by intrinsic long-term variations and/or fluctuations (with which we mean an erratic behavior of the phase delays possibly caused by variations in the instantaneous accretion torques or movements of the accretion footprints on the NS surface, see \\citealp{DiSalvo_07}\\ for a review). Examples of this complex behavior of the pulse phase shifts in accreting millisecond pulsars were already reported in literature. \\cite{Burderi_06}, who analyzed the 2002 outburst of the accreting millisecond pulsar \\saxj\\ found a jump of 0.2 in the pulse phases of the first harmonic which is not present in the second harmonic phases, which show a much more regular behavior. This change is in correspondence of a change in the slope in the exponential decay of the X-ray light curve (see also \\cite{Hartman_07} for a discussion of the complex phase behavior in other outbursts of \\saxj). \\cite{Papitto_07} found that the second harmonic of XTE~J1814--338 follows the first harmonic giving approximately the same spin frequency derivative. A clear model which can explain this behavior is still missing, but these observational evidences seem to suggest that perhaps the second harmonic has a more fundamental physical meaning. For instance it may be related to the emission of both the polar caps while the first harmonic may be dominated by the most intense but less stable polar cap. Another possible explanation comes from possible shape and/or size variations of the accretion footprints related to variations of the accretion rate. \\cite{Romanova_03} found a such behavior in their numerical simulations. In this paper we report the results of a timing analysis performed on \\xtejb, making use of an improved orbital solution \\citep{Riggio_07}. As in the cases mentioned above, \\xtejb\\ shows erratic fluctuations of the phase delays of the first harmonic and a much more regular behavior of the phase delays derived from the second harmonic. In the hypothesis that the second harmonic pulse phase delays are a good tracer of the spin frequency evolution we can derive a spin-up rate of $2.5(7) \\times 10^{-14}$ Hz/s (1 $\\sigma$ confidence level). ", "conclusions": "We have analyzed a long RXTE observation of the accreting millisecond pulsar \\xtejb\\ and reported the results of an accurate timing analysis performed on a time span of about $120$ days, the longest outburst of an accreting millisecond pulsar for which a timing analysis has been performed to date. We find that the phase delays derived from the first harmonic show an erratic behavior around a global parabolic spin-up trend. This behavior is similar to that previously shown by two accreting millisecond pulsar, \\saxj\\ \\citep{Burderi_06} and \\xtejc\\ \\citep{Papitto_07}. In the case of the 2002 outburst of \\saxj, the phase delays of the first harmonic show a shift by about 0.2 in phase at day 14 from the beginning of the outburst, when the X-ray flux abruptly changed the slope of the exponential decay. On the other hand, the phase delays of the second harmonic in \\saxj\\ showed no sign of the phase shift of the first harmonic, and could be fitted by a spin-up in the first part of the outburst plus a barely significant spin-down at the end of the outburst. In the case of \\xtejc, the fluctuations in the phase delays were visible both in the first harmonic and in the second harmonic, superposed to a global parabolic spin-down trend. \\cite{Papitto_07} have shown that the post-fit phase residuals were strongly anti-correlated to variations of the X-ray light curve. These fluctuations were interpreted as due to movements of the accretion footprints (or accretion column) induced by variations of the X-ray flux. In the case of \\xtejb, the fluctuations in the phase delays affect mostly the first harmonic, which shows a trend that is very difficult to reproduce with a simple model. As in the case of \\xtejc, the post-fit phase residuals are clearly anti-correlated with variations observed in the X-ray light curve; from Figure~\\ref{fig1} we see that the phases decrease when the X-ray flux shows rapid increases. It is important to note that the anti-correlation visible between the post-fit phase delays and the X-ray flux is independent of the spin-down or spin-up behavior of the source, since it is observed in \\xtejc, which shows spin-down, and in \\xtejb, which shows spin-up. The correlation between the phase delays and the X-ray flux affects the second harmonic only marginally. Indeed, there are a few points in the phase delays of the second harmonic that are significantly below the global trend observed in the phase delays, and all of them correspond to flares in the X-ray light curve. Excluding these points marginally affects the values we obtain for the spin frequency and its derivative, but gives a significant improvement of the $\\chi^2$ of the fit. We find that the phase delays of the harmonic are fitted by a parabolic spin-up model. We have also showed that the quality of the fit is much improved if we use a more physical model in which the spin-up rate decreases exponentially with time following the decrease of the X-ray flux (and hence of the inferred mass accretion rate). In fact, if the spin-up of the source is related to the mass accretion rate, then it should not be constant with time, but, in first approximation, should decrease proportionally with the mass accretion rate onto the NS. For instance, assuming that accretion of matter and angular momentum occurs at the corotation radius $R_{co}$, the relation between the spin frequency derivative and the mass accretion rate is, from the angular momentum conservation law, $\\nudot = \\Mdot \\, \\sqrt{G M R_{co}}/ 2 \\pi I$, where G is the gravitational constant, M the NS mass and I is the NS moment of inertia; this gives a lower limit on the mass accretion rate since the specific angular momentum at the corotation radius is the maximum that can be transferred to the NS. In the case of \\xtejb, the duration of the outburst is particularly long (about 120 days), and the effect of the global decrease of the mass accretion rate during the outburst should be particularly relevant for this source. Indeed in this case the fit we obtain using an exponentially decreasing spin-up rate is significantly better than using a constant spin-up rate. From the fit of the phase delays of the second harmonic of \\xtejb\\ with the model discussed above we find a mass accretion rate at the beginning of the outburst of $4(1) \\times 10^{-10}$ M$_\\odot$ yr$^{-1}$.\\footnote{For this estimation we adopted the value of $I=10^{45}$ g cm$^2$, M = 1.4 M$_\\odot$ and NS radius $R_{NS} = 10^6$ cm.} We can compare this mass accretion rate with the X-ray flux of the source at the beginning of the outburst that was $2 \\times 10^{-9}$ ergs cm$^{-2}$ s$^{-1}$ \\citep{Falanga_05} from which we derive an X-ray luminosity of $4.7 \\times 10^{36}$ ergs s$^{-1}$ and a distance to the source of 4.4(6) kpc. Clearly this is only a crude estimation of the distance on the basis of our timing results and future independent estimates are needed to confirm or disprove our hypothesis." }, "0710/0710.1663_arXiv.txt": { "abstract": "Recent observations by \\xmm\\ detected rotational pulsations in the total brightness and spectrum of several neutron stars. To properly interpret the data, accurate modeling of neutron star emission is necessary. Detailed analysis of the shape and strength of the rotational variations allows a measurement of the surface composition and magnetic field, as well as constrains the nuclear equation of state. We discuss our models of the spectra and light curves of two of the most observed neutron stars, \\rxj\\ and \\onee, and discuss some implications of our results and the direction of future work. ", "introduction": "} Thermal radiation from the surface of neutron stars (NSs) can provide invaluable information on the physical properties and evolution of NSs. NS properties, such as the mass $M$ and radius $R$, in turn depend on the poorly constrained physics of the stellar interior, such as the nuclear equation of state (EOS) and quark and superfluid/superconducting properties at supra-nuclear densities. Many NSs are also known to possess strong magnetic fields ($B \\sim 10^{12}-10^{13}$~G), with some well above the quantum critical value ($B\\gg B_{\\rm Q}\\equiv 4.4\\times 10^{13}$~G). The observed thermal radiation originates in a thin atmospheric layer (with scale height $\\sim 1$~cm) that covers the stellar surface. To properly interpret the observations of NS surface emission and to provide accurate constraints on their physical properties, it is important to understand in detail the radiative behavior of NS atmospheres in the presence of strong magnetic fields (see \\citep{pavlovetal95,holai01,zavlinpavlov02,holai03,zavlin07}, for more detailed references on observations and on previous works in NS atmosphere modeling). The properties of the atmosphere, such as the chemical composition, EOS, and radiative opacities, directly determine the characteristics of the observed spectrum. While the surface composition of the NS is unknown, a great simplification arises due to the efficient gravitational separation of light and heavy elements \\citep{alcockillarionov80}. A pure hydrogen atmosphere is expected even if a small amount of accretion/fallback occurs after NS formation; the total mass of hydrogen needed to form an optically thick atmosphere can be less than $\\sim 10^{16}$~g. On the other hand, a heavy element atmosphere may be possible if no accretion takes place. The strong magnetic fields present in NS atmospheres significantly increase the binding energies of atoms, molecules, and other bound states (see \\citep{lai01}, for a review). Abundances of these bound states can be appreciable in the atmospheres of cold NSs (i.e., those with surface temperature $T\\lesssim 10^6$~K; \\citep{laisalpeter97,potekhinetal99}). In addition, the presence of a magnetic field causes emission to be anisotropic and polarized; this must be taken into account when developing radiative transfer codes. The most comprehensive early studies of magnetic NS atmospheres focused on a fully ionized hydrogen plasma and moderate field strengths ($B\\sim 10^{12}-10^{13}$~G; \\citep{miller92,shibanovetal92,pavlovetal94,zaneetal00}). These models are expected to be valid only for relatively high temperatures. More recently, atmosphere models in the ultra-strong field ($B\\gtrsim 10^{14}$~G) and relevant temperature regimes have been presented (\\citep{ozel01,zaneetal01,holai03,lloyd03,vanadelsberglai06}; see also \\citep{bezchastnovetal96,bulikmiller97}, for early work), and all of these rely on the assumption of a fully ionized hydrogen composition. Magnetized non-hydrogen atmospheres have been studied by \\citep{miller92,rajagopaletal97}, but because of the complexity of the atomic physics, the models were necessarily crude (see \\citep{moriho07}, for more details). Only recently has self-consistent atmosphere models \\citep{hoetal03,potekhinetal04,moriho07} using the latest EOS and opacities for partially ionized hydrogen \\citep{potekhinchabrier03,potekhinchabrier04} and mid-$Z$ elements \\citep{morihailey02,morihailey06} been constructed. The atmosphere models discussed above only describe emission from a local patch of the stellar surface. By taking into account surface magnetic field $\\vecB$ and temperature $T$ distributions, we can construct more physically correct models of emission from NSs. However, these spectra from the whole NS surface are necessarily model-dependent, as the $\\vecB$ and $T$ distributions are unknown. Nevertheless, detailed comparisons of the models with rotation phase-resolved observations is a powerful tool to study NSs, e.g., spectral features that vary with phase are essential to disentangling magnetic field effects from other parameters and to probe the magnetic field geometry on the surface of the star. Indeed there have been recent works attempting to fit magnetic atmosphere spectra to observations of NSs (see \\citep{ho07}, and references therein). Here we describe some of the details and observational applications of our work. ", "conclusions": "" }, "0710/0710.2899_arXiv.txt": { "abstract": "We report the detection of \\hi~21~cm absorption from the $z=2.289$ damped Lyman-$\\alpha$ system (DLA) towards TXS~0311+430, with the Green Bank Telescope. The 21~cm absorption has a velocity spread (between nulls) of $\\sim 110$~km~s$^{-1}$ and an integrated optical depth of $\\int \\tau {\\rm d}V = (0.818 \\pm 0.085)$~km~s$^{-1}$. We also present new Giant Metrewave Radio Telescope 602~MHz imaging of the radio continuum. TXS~0311+430 is unresolved at this frequency, indicating that the covering factor of the DLA is likely to be high. Combining the integrated optical depth with the DLA \\hi~column density of \\nhi\\ = $(2 \\pm 0.5) \\times 10^{20}$~\\cm, yields a spin temperature of $\\ts = (138 \\pm 36)$~K, assuming a covering factor of unity. This is the first case of a low spin temperature ($< 350$~K) in a $z > 1$ DLA and is among the lowest ever measured in any DLA. Indeed, the $\\ts$ measured for this DLA is similar to values measured in the Milky Way and local disk galaxies. We also determine a lower limit (Si/H)~$\\gtrsim 1/3$ solar for the DLA metallicity, amongst the highest abundances measured in DLAs at any redshift. Based on low redshift correlations, the low $\\ts$, large 21~cm absorption width and high metallicity all suggest that the $z \\sim 2.289$ DLA is likely to arise in a massive, luminous disk galaxy. ", "introduction": "\\label{sec:intro} Damped Lyman-$\\alpha$ systems (DLAs) are the highest column density absorbers seen along QSO lines-of-sight, with neutral hydrogen column densities \\nhi\\ $\\ge 2.0\\times 10^{20}$~\\cm. They have long been identified as the precursors of today's galaxies and the primary gas reservoir for star formation at high redshifts. Despite the recognised importance of the absorbers, their typical size, structure and internal physical conditions remain issues of controversy (e.g. \\citealt{wolfe05}). For example, DLA metallicities, now measured in over 100 absorbers, show very little evolution between $z \\sim 3$ and $z \\sim 0$, with low-metallicity ($-2<$~[Z/H]~$< -1$, where Z~$\\equiv$~Zn, S or Si\\footnote{In the usual notation, [Z/H]~$\\equiv$~log(N(Z)/N(\\hi))$-$log(N(Z)/N (\\hi))$_{\\odot}$. We use solar values from \\citet{lodders03}.}) absorbers the norm at all redshifts (e.g. \\citealt{kulkarni05}). This lack of metallicity evolution runs contrary to expectations that the interstellar metallicity should rise towards lower redshifts if DLAs trace the bulk of gas in galaxies. Moreover, only a few tens of DLAs have their galactic counterparts identified, of which only a small fraction have been spectroscopically confirmed (e.g. \\citealt{chen03}). Our understanding of the basic physical properties of the absorbing galaxies, such as their mass, size, star formation rates and luminosity, remains very limited. \\hi~21~cm absorption studies of DLAs towards radio-loud quasars provide an independent probe of physical conditions in the absorbers. They can be combined with optical measurements of the \\hi\\ column density of the DLA (from the Lyman-$\\alpha$ line) to obtain the column-density-weighted harmonic mean spin temperature $\\ts$ of the absorbing gas, allowing one to determine the temperature distribution of the \\hi\\ along the line of sight. Measurements of $\\ts$ in a large number of DLAs may ultimately be used to infer other galactic properties of high redshift systems. For example, in the local Universe, large spiral disks like the Milky Way and M31 typically have low $\\ts$ values ($\\lesssim 300$~K; \\citealt{braun92}) while high $\\ts$ values ($\\sim 1000$~K; \\citealt{young97}) are more common in dwarf galaxies. Tentative evidence for a similar trend has been found in low~$z$ DLAs, out to $z \\sim 0.7$ \\citep{kanekar03}. If such correlations hold for DLAs at all redshifts, measurements of $\\ts$, and its evolution, would provide interesting insights into galaxy evolution. Unfortunately, despite a number of searches over the past two and a half decades (e.g. \\citealt{briggs83,carilli96,chengalur00,kanekar03}), spin temperature estimates are available for only $15$~DLAs (of which $10$ are lower limits) at $z \\gtrsim 1.7$ (Kanekar et al, in preparation). All of these previous high $z$ measurements yield high spin temperatures of $\\ts \\gg 300$ K. There are a number of reasons why the crop of $\\ts$ estimates has increased slowly over the last 25 years, including observational issues such as the frequency coverage of radio telecopes and widespread radio frequency interference (RFI) at the low frequencies ($\\lesssim 1$~GHz) of the redshifted 21~cm line. However, an additional important reason for the relatively-small present 21~cm absorption sample is simply the dearth of known DLAs towards radio-loud QSOs suitable for 21~cm absorption follow-up. We have hence been conducting an optical survey of low-frequency-selected radio-loud quasars, specifically designed to increase the number of $\\ts$ estimates in the redshift range $2 500$~K, have been detected in 21~cm absorption (of which only four are at $z \\gtrsim 2$; \\citealt{wolfe79,wolfe81,kanekar06,kanekar07}) while the remaining DLAs without detectable 21~cm absorption all have $3\\sigma$ lower limits of $> 700$~K on the spin temperature. The preponderance of high spin temperature estimates in high-$z$ DLAs has usually been attributed to a relatively low fraction of the cold neutral phase of \\hi\\ (the CNM), with most of the gas in the warm phase (the WNM) (e.g. \\citealt{carilli96,chengalur00}). This assumes that the absorbers have a similar two-phase structure to that seen in the ISM of the Milky Way (e.g. \\citealt{wolfe03b}). The low spin temperature of the $z \\sim 2.289$ DLA towards TXS~0311+430 would then imply that a higher fraction of the \\hi\\ along the line of sight is in the CNM compared with the majority of DLAs at $z \\gtrsim 2$. \\citet{chengalur00} argued that the high derived $\\ts$ values of high-$z$ DLAs could be due to the low metallicities of typical high-$z$ DLAs, with the paucity of metals resulting in fewer radiation pathways for gas cooling (see also \\citealt{young97}). If this is correct, one would expect DLAs with low spin temperatures (such as the absorber towards TXS~0311+430) to have significantly higher metallicities than those of the general DLA population \\citep{kanekar01a}. As expected, and assuming that the Si\\,{\\sc ii}~$\\lambda 1808$ equivalent width has not been over-estimated, the $z \\sim 2.289$ DLA has [Si/H]~$\\ge -0.48$, one of most metal-rich DLAs yet discovered. It has also been suggested that the observed low 21~cm optical depth in high-$z$ DLAs arises due to covering factor effects (e.g. \\citealt{curran05}). In the present case, any reduction in the covering factor below the assumed value of unity can only strengthen the case for a low spin temperature (since $\\ts = (138 \\pm 36) \\times f$~K). The only way to alter the conclusion that a sizeable fraction of \\hi\\ along the line of sight is in the cold phase is if there are large spatial differences in the \\hi\\ column densities along the optical and radio lines of sight (e.g. \\citealt{wolfe03b}). For example, if the optical QSO lies behind a ``hole'' in the \\hi\\ column density distribution, the average \\hi\\ column density against the radio QSO could be larger than that measured from the Lyman-$\\alpha$ line, implying a higher spin temperature from Eqn.~\\ref{eqn:tspin}. Unfortunately, it is very difficult to directly test this possibility. While Galactic \\hi\\ column densities derived from Lyman-$\\alpha$ absorption studies are in excellent agreement with those obtained from \\hi\\ 21~cm emission observations in the same directions, despite the very different spatial resolutions in the two methods, such comparisons have only been carried out for a fairly small number of high latitude lines of sight \\citep{dickey90}. A similar comparison between \\nhi\\ values derived from Lyman-$\\alpha$ absorption and 21~cm emission has only been possible in one DLA, the $z \\sim 0.009$ absorber towards SBS~1543+593. Here, the \\hi\\ column densities agree to within a factor of $\\sim 2$, despite the extremely poor spatial resolution ($\\sim 5.3 \\: h_{71}^{-1}$~kpc) of the radio observations \\citep{chengalur02}. While both of these studies suggest that the \\nhi\\ values along the radio and optical lines of sight are likely to be comparable, we cannot formally rule out the possibility of differences in an individual absorber. However, the fact that the expected high metallicity is indeed seen in the $z \\sim 2.289$ DLA (assuming that our Si~{\\sc ii}~$\\lambda$1808 measurement is accurate) is consistent with the interpretation of a high CNM fraction. Finally, \\citet{kanekar03} noted a relationship between spin temperature and absorber morphology, in that, at low redshifts, low $\\ts$ values are only found in DLAs identified with luminous disk galaxies, while low-$z$, low-luminosity DLAs, associated with dwarf or LSB galaxies, are all found to have high spin temperatures ($\\ts \\gtrsim 700$~K). It has not been possible to test this empirical relationship at $z > 1$, as the host galaxies of high-$z$ DLAs are rarely detectable. However, if the relationship does extend out to high redshifts, we would expect the $z \\sim 2.289$ DLA to be a massive, luminous disk galaxy. We note that both the high metallicity and the large velocity spreads seen in the optical low-ionization metal lines and the 21cm absorption are consistent with the absorption arising in a massive galaxy. If so, it should be possible to detect the absorber host with deep imaging; this would be the first direct test of the $\\ts$-morphology relationship at high redshifts. In summary, we have detected damped Lyman-$\\alpha$ and \\hi~21~cm absorption at $z = 2.289$ towards the quasar TXS~0311+430. We obtain a DLA spin temperature of $\\ts = (138 \\pm 36) \\times f$~K, the first case of a low spin temperature estimate in a high redshift DLA. The low spin temperature, high metallicity and large velocity spread of the 21~cm and metal lines all suggest that the absorber is likely to be a massive disk galaxy." }, "0710/0710.0252_arXiv.txt": { "abstract": "{The ANTARES telescope is being built in the Mediterranean Sea. The detector consists of a 3D array of photomultipliers (PMTs) that detects the Cherenkov light induced by the muons produced in neutrino interactions. Other signatures can also be detected. Since the neutrino fluxes from point-like sources are expected to be small, it is of the utmost importance to take advantage of the ANTARES pointing accuracy (angular resolution better than 0.3 degrees for muon events above 10 TeV) to disentangle a possible signal from the unavoidable atmospheric neutrino background. In order to distinguish an excess of neutrino events from the background, several searching algorithms have been developed within the ANTARES collaboration. In this contribution, the discovery potential and sensitivity to point-like sources of the ANTARES neutrino telescope are presented.} \\begin{document} ", "introduction": "The ANTARES collaboration~\\cite{ANTARES} has started the construction of an underwater neutrino telescope in the Mediterranean Sea at a depth of 2475~m. Seven of the twelve lines of the detector have been deployed so far (May 2007) and the whole detector will be commissioned in early 2008. Each line is equipped with 75 10'' photomultiplier tubes (PMTs) joined in triplets making 25 floors along the line. The lines are kept vertical by means of a buoy located at their upper end. The mean distance among lines is about 65 m and the instrumented length starts at 100 m over the seabed and covers about 350 m. The angular resolution for the ANTARES telescope at high energies (above 10~TeV) is better than 0.3$^{\\circ}$. At lower energies, the angular resolution is dominated by the angle between the muon track and the original neutrino direction. The angular resolution together with the effective area determine the performance of a neutrino telescope. The good pointing accuracy and large effective area at high energies (0.06 km$^2$ at 100 TeV) enables us to search for point-like neutrino sources or, in case no hint of neutrinos source is found, to set restrictive upper limits to neutrino fluxes. ", "conclusions": "In this contribution we have reviewed the expected performance of the ANTARES neutrino telescope regarding the search for point-like neutrino sources. Several searching algorithms have been devised in ANTARES. Among them, unbinned techniques turn out to be more efficient than the standardised binning approach. In addition, the method based on the EM algorithm presents very good results without the trade-off of higher dependence on the estimated detector performances. % Discovery potentials in terms of number of events and required neutrino flux to claim the existence of a source at 50\\% of probability has been presented. The comparison with other experiments in term of the sensitivity has also been presented. \\\\ \\\\ {\\it This work is supported by Spanish MEC grants FPA2003-00531 and FPA2006-04277.}" }, "0710/0710.2127_arXiv.txt": { "abstract": "We have used high precision differential astrometry from the Palomar High-precision Astrometric Search for Exoplanet Systems (PHASES) project and radial velocity measurements covering a time-span of 20 years to determine the orbital parameters of the 88 Tau A system. 88 Tau is a complex hierarchical multiple system comprising a total of six stars; we have studied the brightest 4, consisting of two short-period pairs orbiting each other with an $\\sim$18-year period. We present the first orbital solution for one of the short-period pairs, and determine the masses of the components and distance to the system to the level of a few percent. In addition, our astrometric measurements allow us to make the first determination of the mutual inclinations of the orbits. We find that the sub-systems are not coplanar. ", "introduction": "88 Tau (HD 29140, HR 1458, HIP 21402) is a bright ($m_V=4.25, m_K=3.69\\pm 0.25$; Skrutskie et al. 2006\\nocite{2mass}), nearby ($\\sim 50$pc) hierarchical sextuple stellar system \\citep{msc}. The A component contains a pair of systems (designated Aa and Ab) in an $\\sim18$-year \\citep{balega99} orbit that has been resolved by speckle interferometry \\citep{mac87}. The Aa component is a known spectroscopic binary system ($P\\sim 3.57$-day), with a composite spectral type of A5m \\citep{c69}. In previous work it had been noted \\citep{bc88} that the A system is likely complex, with possibly as many as 5 components. \\citet{balega99} noted a discrepancy between the total estimated mass of this system based on photometry and spectral types, and the total mass derived from the visual orbit and {\\em Hipparcos} parallax. In this work we have determined that, like the Aa component, the Ab component is a double-lined binary; this newly-resolved binary has a period of 7.89 days. Finally, there is a common-proper-motion companion, labeled B, located $\\sim$69 arcseconds away from the A system; it, too, is known to be a binary \\citep{tg01}. For clarity we provide a schematic of this complex system in Figure \\ref{fig1}. \\begin{figure} \\figurenum{1} \\plotone{f1.eps} \\figcaption{A schematic diagram of the 88 Tau system. \\label{fig1}} \\end{figure} There are several reasons why multiple stellar systems such as 88 Tau merit attention: first, binary orbits make it possible to measure accurate stellar masses and distances, while the larger number of presumably co-eval stars allows one to impose the additional constraint that any given model must accurately match all of the stars. This approach has proven particularly fruitful when applied to another famous hierarchical sextuple system: Castor\\citep{tr02}. Second, as outlined in \\citet{st02}, the relative orientations of the orbital angular momenta allow one to constrain the properties of the cloud from which the stars are thought to have formed, as well as the subsequent dynamical decay process. Despite their value, observational problems have limited the number of triple or higher-order systems with accurately measured orbits to fewer than 10. Given their hierarchical nature it is often the case that either the close system is unresolvable or the outer system has an impractically long orbital period. With the advent of long-baseline stellar interferometry, and more recently phase-referenced long-baseline interferometric astrometry \\citep{lm04} capable of 10--20 $\\mu$-arcsecond astrometric precision between pairs of stars with separations in the range 0.05--1 arcsecond, it has become possible to resolve the orbital motion of several interesting multiple systems \\citep{kapPeg,v819Her}. Here we report on astrometric and radial velocity measurements of the 88 Tau A system, which allow us to constrain the orbits of the 3.57-day, 7.89-day and 18-year components with improved precision, and for the first time provide a relative orientation of the orbits as well as component masses. Astrometric measurements were made with the Palomar Testbed Interferometer \\citep{colavita99} as part of the Palomar High-precision Astrometric Search for Exoplanet Systems (PHASES) program \\citep{limits}. The Palomar Testbed Interferometer is located on Palomar Mountain near San Diego, CA. It was developed by the Jet Propulsion Laboratory, California Institute of Technology for NASA as a testbed for interferometric techniques applicable to the Keck Interferometer and the Space Interferometry Mission (SIM). It operates in the J (1.2 $\\mu$m), H (1.6 $\\mu$m), and K (2.2 $\\mu$m) bands and combines starlight from two out of three available 40 cm apertures. The apertures form a triangle with 86 and 110 m baselines. ", "conclusions": "PHASES interferometric astrometry has been used together with radial velocity data to measure the orbital parameters of the quadruple star system 88 Tau A, and in particular to resolve the apparent orbital motion of the close Aa1-Aa2 and Ab1-Ab2 pairs. We have made the first determination of the period of the Ab binary system and found it to consist of a pair of nearly equal-mass G stars. The amplitude of the Ab1-Ab2 Center-of-Light motion is only $\\sim 65 \\mu$as, indicating the level of astrometric precision attainable with interferometric astrometry. We are able to resolve the orbital motion of all of the components, and hence determine the orbital inclinations and component masses with a precision of a few percent. Finally, we are able to determine the mutual inclinations of the various orbits." }, "0710/0710.5531_arXiv.txt": { "abstract": "{} {Procyon A, a bright F5 IV-V Sun-like star, is justifiably regarded as a prime asteroseismological target. This star was repeatedly observed by MOST, a specialized microsatellite providing long-term, non-interrupted broadband photometry of bright targets. So far, the widely anticipated p modes eluded direct photometric detection, though numerous independent approaches hinted for the presence of signals in the f$\\sim0.5-1.5$\\,mHz range.} {Implementation of an alternative approach in data processing, as well as combination of the MOST data from 2004 and 2005 (264\\,189 measurements in total) helps to reduce the instrumental noise affecting previous reductions, bringing the $3\\sigma$ detection limit down to $\\sim$5.5\\,part-per-million in the $f=0.8-1.2$\\,mHz range.} {This enables to cross-identifiy 16 p-mode frequencies (though not their degrees) which were previously detected via high-precision radial velocity measurements, and provides an estimate of the large spacing, $\\delta\\nu =0.0540$\\,mHz at $f\\sim1$\\,mHz. The relatively low average amplitude of the detected modes, $a=5.8\\pm0.6$\\,ppm, closely matches the amplitudes inferred from the ground-based spectroscopy and upper limits projected from WIRE photometry. This also explains why such low-amplitude signals eluded the direct-detection approach which exclusively relied on the MOST 2004 (or 2005) data processed by a standard pipeline.} {} ", "introduction": "Procyon A (F5 IV-V) has served as a prime target in many asteroseismological campaigns (see the reviews by Bedding \\& Kjeldsen, 2006,2007, and references therein). While high-precision measurements of radial velocities revealed the presence of variability caused by solar-like oscillations (\\cite{ma99}) and provided a preliminary identification of p-mode frequencies (\\cite{egge}; \\cite{mart}; \\cite{lecc}), the photometric studies met either limited success (\\cite{brun}) or outright null-detection (\\cite{matt}; \\cite{gun}). The null-detection result was hotly debated ever since (\\cite{bedd}; \\cite{regu}). Attempting to resolve the controversy, we applied an alternative data-processing approach to the abundant photometric data acquired by MOST in 2004-2005. ", "conclusions": "" }, "0710/0710.5707_arXiv.txt": { "abstract": "In February 1997, the Japanese radio astronomy satellite HALCA was launched to provide the space-bourne element for the VLBI Space Observatory Programme (VSOP) mission. Approximately twenty-five percent of the mission time was dedicated to the VSOP Survey of bright compact Active Galactic Nuclei (AGN) at 5~GHz. This paper, the fifth in the series, presents images and models for the remaining 140 sources not included in Paper III, which contained 102 sources. For most sources, the plots of the \\uv\\ coverage, the visibility amplitude versus \\uv\\ distance, and the high resolution image are presented. Model fit parameters to the major radio components are determined, and the brightness temperature of the core component for each source is calculated. The brightness temperature distributions for all of the sources in the VSOP AGN survey are discussed. ", "introduction": "The radio astronomy satellite HALCA (Highly Advanced Laboratory for Communications and Astronomy) was launched by the Institute of Space and Astronautical Science in February 1997 to participate in Very Long Baseline Interferometry (VLBI) observations with arrays of ground radio telescopes. HALCA provides the longest baselines of the VSOP, an international endeavor that has involved over 28 ground radio telescopes, five tracking stations and three correlators \\citep{hir98,hir00a}. HALCA was placed in an orbit with an apogee height above the Earth's surface of 21,400\\,km, a perigee height of 560\\,km, and an orbital period of 6.3~hours. During the seven years of HALCA's mission lifetime, about 75\\% of observing time was used for projects selected by international peer-review from open proposals submitted by the astronomical community in response to Announcements of Opportunity. This part of the mission's scientific programme constituted the General Observing Time (GOT). The remaining observing time was devoted to a mission-led survey of active galactic nuclei at 5\\,GHz: the VSOP Survey Program. The major goal of the Survey was to determine the statistical properties of the sub-milliarcsecond structure of a complete sample of AGNs. \\citep{hir00b, fom00a}. % Following the end of the formal international mission period in February 2002, the Japanese-dominated effort continued survey observations until October 2003, when HALCA lost its attitude control capability. This occurred well after the end of the original planned mission lifetime. This paper is the fifth in the series of VSOP Survey related papers. \\citet{sco04} (henceforth P-III) contains the results for 102 sources which were observed and reduced before 2001 October. \\citet{hor04} (henceforth P-IV) analyzed the cumulative visibilities of those sources to obtain the `typical source structure'. This paper contains the additional 140 survey sources which were successfully observed by VSOP and completes the survey programme observing results. The brightness temperature properties of the entire sample of sources are discussed. ", "conclusions": "We have presented images, models and comments of the 140 sources which were observed as part of the VSOP Survey project that were not covered in P-III. We have combined the brightness temperature measurements and limits found for the entire sample to produce the T$_b$ distribution for the VSS. We find that about half of the AGN sample of sources reported upon in this paper have significant radio emission in the core component, with $T_b \\ge 10^{12}$~K in the source frame. Since the maximum brightness temperature one is able to determine using only ground-based arrays is of the order of $10^{12}$~K, our results confirm the necessity of using space VLBI to explore the extremely high brightness temperature regime. In addition, our Survey results clearly show that by using space VLBI with higher sensitivity, and somewhat higher resolution, the radio cores of many AGN can be successfully imaged. Because of the variability of many of the sources in the Survey sample, detailed spectral indices of the core components are difficult to determine. However, many of the sources were observed with the VLBA at 15 GHz as part of the VLBA2cm2 survey, and the spectral properties of the cores will be reported elsewhere \\citep{VLBA2cm2}. It was not possible to slew the HALCA satellite during the observing runs, therefore HALCA was not able to participate in scans of fringe finders, or flux calibrators. It is the absence of these which forces us to label a number of experiments with no space fringes as failures, when it could be the effects of the source structure. The design of VSOP-2 will allow fringe checks, and also phase referencing experiments, to be performed \\citep{hir00c}. {% The completion of this survey has been a Quixotic endeavor, and possibly: ``vino a dar en el m\\'as estra\\~no pensamiento que jam\\'as dio loco en el mundo; y fue que le pareci\\'o convenible y necesario, as\\'i para el aumento de su honra como para el servicio de su rep\\'ublica, hacerse caballero andante'' \\citep{quixote}. }" }, "0710/0710.3121_arXiv.txt": { "abstract": "We have examined images from the Precision Solar Photometric Telescope (PSPT) at the Mauna Loa Solar Observatory (MLSO) in search of latitudinal variation in the solar photospheric intensity. Along with the expected brightening of the solar activity belts, we have found a weak enhancement of the mean continuum intensity at polar latitudes (continuum intensity enhancement $\\sim0.1 - 0.2\\%$ corresponding to a brightness temperature enhancement of $\\sim2.5{\\rm K}$). This appears to be thermal in origin and not due to a polar accumulation of weak magnetic elements, with both the continuum and CaIIK intensity distributions shifted towards higher values with little change in shape from their mid-latitude distributions. Since the enhancement is of low spatial frequency and of very small amplitude it is difficult to separate from systematic instrumental and processing errors. We provide a thorough discussion of these and conclude that the measurement captures real solar latitudinal intensity variations. ", "introduction": "Latitudinal variations in the thermal structure of the solar convective envelope have been implied or required by theories ranging from non-general-relativistic gravity \\citep[]{dic64} to those explaining the solar differential rotation \\citep[e.g.][]{kit95,dur99,rem05,mie06}, meridional circulation, and torsional oscillations \\citep[]{spr03}. Models of the solar differential rotation and meridional circulation reproduce the observed helioseismic rotation profiles \\citep[]{thom96,sch98} best when baroclinicity is introduced via a latitudinal temperature gradient in region of the solar tachocline which spreads into the convection zone proper by turbulent convection \\citep[]{rem05,mie06}. While it is estimated from such models that in the solar photosphere the pole may be as much as a few degrees warmer and the mid-latitudes a couple of degrees cooler than the equatorial region \\citep[e.g][]{bru02,mie06}, these estimates are model dependent and measurements have proven difficult. The difficulties arise because of the intrinsically low amplitude thermal signal expected and the presence of magnetic structures which introduce small scale intensity fluctuations of comparable or greater amplitude. Thus, previous observations have yielded wide ranging results (Table~\\ref{table1}). Here we examine full disk images from the Precision Solar Photometric Telescope (PSPT) at the Mauna Loa Solar Observatory (MLSO). The telescope should by design readily allow measurement of latitudinal temperature variations if they are a few degrees in magnitude. In \\S2 below we review the telescope capabilities and data processing techniques necessary to realize those. In \\S3 we present the results obtained and in \\S4 discuss the uncertainties in these due to random and systematic errors. Finally, we conclude in \\S5 by putting our methods and results in the context of previous measurements. ", "conclusions": "We have measured an enhancement of the photospheric continuum intensity in the solar polar regions using PSPT images. We have carefully examined the properties of the signal and believe it to be of solar origin. Moreover, the properties of the intensity distributions suggest that the signal is thermal not magnetic in origin, although contributions from very weak network elements associated with the pole-ward extending branch of the solar cycle \\citep[e.g.][]{how81,mul84,she06} cannot be ruled out without processing significantly more data and employing masking of greater severity. This is planned, but limitations of the current ground based solar photometric observations may make a decisive measurement difficult. Space based imaging photometry would be ideal for this and other intriguing problems. This is particularly true if solar cycle dependencies are to be uncovered. Such measurements would help constrain global solar models by adding a thermodynamic constraint to the dynamical constraints currently provided by helioseismology \\citep[]{kld88}. Current numerical simulations seem to show little indication of a local temperature minimum at the equator, as our surface measurement implies, but rather display a mid-latitude minimum with a weak local maximum at the equator and stronger one at the pole. This is true even when a monotonically decreasing pole to equator profile is imposed in the solar tachocline \\citep[]{rem05,mie06}. While we find some hint of equatorial brightening after application of our most restrictive activity masks (bottom row Figure~\\ref{fig3}), it has extremely low amplitude and the measurement is extremely uncertain. Our efforts can be distinquished from previous work in several ways. Early work (references $1-7$ in Table~1) struggled with instrumental precision, systematic error, and detector gain correction. We too have taken pains to address these. More importantly, no distinction was made in that early work between magnetic and thermal contributions to the signal, other than perhaps qualitative selection of nonactive regions. We have emphasized this distinction by applying a series of masks based on CaIIK intensity to the images before analysis. Previous efforts by Kuhn et al. (references $8-10$ in Table~$10$) have, for the most part, also attempted to separate quiet Sun and facular contributions to the latitudinal signal. The approach in those studies differed from ours in that the facular contribution to the latitudinal variation was estimated statistically based on models of the quiet Sun and facular color and intensity distributions. Interestingly, the pole/mid-latitiude temperature contrast measured in those studies is in good agreement with our measurement. The previous authors, however, reported an additional low latitude brightness enhancement, only seeen very marginally, if at all, in our study at the most severe masking levels. We note, that the one measurement of solar latitudinal intensity variation available using spacecraft data \\citep[]{kuh98} did not remove the facular contribution during analysis. The latitudinal temperature profile obtained in that work looks very similar to our Figure~$6a$, also unmasked. Finally, we caution that our analysis does not completely eliminate possible contributions to the solar photospheric intensity from an abundance of unresolved magnetic elements of very low magnetic flux density. Correlations between magnetic flux density and CaIIK intensity extend to very low values, making masking based on these images possible, but also suggesting continued contamination of the continuum measurements at very low contrasts (Rast 2003). Unambiguous measurement of a strictly thermal signal in the solar photosphere may require space-based full-disk photometric imaging at high precision and resolution over many wavelengths, combined with detailed modeling of the radiation field. Even that effort may be stymied if in fact there is no quiet Sun." }, "0710/0710.4537_arXiv.txt": { "abstract": "Due to projection effects, coronagraphic observations cannot uniquely determine parameters relevant to the geoeffectiveness of CMEs, such as the true propagation speed, width, or source location. The Cone Model for Coronal Mass Ejections (CMEs) has been studied in this respect and it could be used to obtain these parameters. There are evidences that some CMEs initiate from a flux-rope topology. It seems that these CMEs should be elongated along the flux-rope axis and the cross section of the cone base should be rather elliptical than circular. In the present paper we applied an asymmetric cone model to get the real space parameters of frontsided halo CMEs (HCMEs) recorded by SOHO/LASCO coronagraphs in 2002. The cone model parameters are generated through a fitting procedure to the projected speeds measured at different position angles on the plane of the sky. We consider models with the apex of the cone located at the center and surface of the Sun. The results are compared to the standard symmetric cone model. ", "introduction": "A halo coronal mass ejection (HCME) was first recorded by Howard in 1982 (Howard \\emph{et al.}, 1982). Since then, HCME are routinely recorded in white light by coronagraphs placed in space. In coronagraphic observations, HCMEs appear as an enhancement surrounding the entire occulting disk. HCMEs originating close to the disk center are often responsible for the severest geomagnetic storms (Gosling, 1993; Kahler, 1992; Webb \\emph{et al.}, 2000). For space weather forecast it is very important to determine the kinetic and geometric parameters describing HCMEs. Unfortunately coronagraphic observations are subjected to projection effects. Viewing in the plane of the sky does not allow to determine the true 3-D space velocity, width and source location of a given CME. It is widely accepted that the geometrical structure of CMEs may be described by the cone model (e.g., Howard \\emph{et al.}, 1982; Fisher and Munro, 1984; St.Cyr \\emph{et al.}, 2000; Webb \\emph{et al}, 2000; Zhao \\emph{et al.}, 2002; Michalek \\emph{et al.}, 2003; Xie \\emph{et al.}, 2004; Xue \\emph{et al.}, 2005). Assuming that the shape of HCMEs is a cone and they propagate with constant angular widths and speeds, at least in their early phase of propagation, a technique was developed (Michalek \\emph{et al.}, 2003) which can determine the following parameters: the linear distance $r$ of the source location measured from the solar disk center, the angular distance $\\gamma$ of the source location measured from the plane of the sky, the angular width $\\alpha$ (cone angle =$0.5\\alpha$) and the 3-D space velocity $V_{S}$ of a given HCME. This technique required measurements of the sky-plane speed and the moment of the first appearance of the halo CME above opposite limbs. If we determine spatial parameters using only two measurements, large random errors would occur. What is more, this technique was limited to asymmetric events not originating from close to the center of the Sun. A similar cone model was used recently by Xie \\emph{et~al.}~(2004) to determine the angular width and orientation of HCMEs. To improve accuracy, in the present attempt the space parameters of HCMEs are determined by fitting the cone model to projected speeds ($V_{P}$) obtained from height-time plots at different position angles. Although many thousands of CMEs were recorded by LASCO coronagraphs the 3D structure of CMEs is still open question. Many authors believe that CMEs originate from a flux-rope geometry, (e.g., Chen \\emph{et al.}, 1997; Dere \\emph{et al.}, 1999; Chen \\emph{et al.}, 2000; Plunket \\emph{et al.}, 2000; Forbes 2000; Krall \\emph{et al.}, 2001; Chen and Krall 2003). If CMEs have a flux-rope geometry, they should be elongated along the flux-rope axis and the cross section of the cone base should be rather elliptical than circular. In the present approach we consider the asymmetric cone model where an eccentricity and orientation of the cone base are new free parameters. We try to identify where to locate the apex of the cone, either at the center of the Sun (Zhao \\emph{et al.}, 2002; Xie \\emph{et al.}, 2004) or on the solar surface (Michalek \\emph{et al.}, 2003). It is important to note that the real elliptical cone model was first developed by Cremades and Bothmer (2004). The model was introduced based on observations of cylindrical shaped CMEs (Cremades and Bothmer, 2004, 2005). They applied it to 32 halo CMEs. In their approach, the best parameter values describing the ellipse are determined from a LASCO image sequence that showed a sharp leading edge. In our method we derive the best-fit parameter values for halo CMEs by working in the velocity space. The paper is organized as follows: in Section~2 the asymmetric cone model is presented, numerical simulations and fitting procedure are explained in Section~3 and in Section~4 the results are described. Final conclusions are given in Section 5. ", "conclusions": "In this paper we have presented the new asymmetric cone model to determine the geometrical parameters characterizing HCMEs. We applied this model to all frontsided HCMEs originating close to the disk center and listed in the SOHO/LASCO catalog in 2002. We estimated, with very good accuracy (the average r.m.s error is equal to 60~km~s$^{-1}$), the cone parameters for $15$ events. It was shown (from the polar plots and the r.m.s errors) that the ACM can reproduce, around the entire occulting disk, the measured projected speed much better than the SCM. We have to note that sometimes the individual projected speeds may significantly differ from the fitting model even for the ACM. In these cases errors are probably caused by inaccurate measurements. Unfortunately, most HCMEs are not sufficiently bright around the entire occulting disk to generate the height-time plots with the same good accuracy for all considered position angles. The biggest errors may arise when the faintest part of a given event is considered or when measurements are disturbed by another event appearing in the LASCO field of view. We also confirmed that most of the events have better fits when the cone apex is located at the center of the Sun. We used this model for the frontsided full HCMEs originating close to the disk center but it can also be applied to limb and even partial HCMEs. When considering limb or partial HCMEs the accuracy will be slightly worse. The determined parameters for HCMEs are similar to that derived by the other cone models, but for some events differences could amount to even 20\\%. We have to remember that this approach has two shortcomings: (1) CMEs may be accelerating, moving with constant speed or decelerating at the beginning phase of propagation. This means that the constant velocity assumption may be invalid.(2) CMEs may expand in addition to radial motion. Then the measured sky-plane speed is a sum of the expansion speed and the projected radial speed. This would also imply that CMEs may not be a rigid cone as we had assumed. Unfortunately, having observation from only one spacecraft there is no possibility to overcome these assumptions. There are two attempts to get the real space parameters of CMEs. Some scientist (e.g. Xie \\emph{et al.}, 2004) try to get the real parameters of CMEs by considering the elliptical shapes of observed CMEs. In our approach we consider velocities of CMEs as a function of the position angle. We take into account the projected velocities (many points) from an entire halo CME, not only from a few chosen points. We do not introduce any assumption about the apparent (geometrical) shapes of CMEs. Therefore, we expect that our determination of parameters may have better accuracy. The determined real velocities of halo CMEs could be potentially useful for space weather forecast. But we have to note that although it is true that faster interplanetary CMEs are more geoeffective, the strength and topology of magnetic field is still crucial to define geoeffectiveness of CMEs (e.g., Bothmer, 2003)." }, "0710/0710.1062_arXiv.txt": { "abstract": "In this paper, we report on the gas-phase abundance of singly-ionized iron (Fe II) for 51 lines of sight, using data from the {\\it Far Ultraviolet Spectroscopic Explorer} ({\\it FUSE}). Fe II column densities are derived by measuring the equivalent widths of several ultraviolet absorption lines and subsequently fitting those to a curve of growth. Our derivation of Fe II column densities and abundances creates the largest sample of iron abundances in moderately- to highly-reddened lines of sight explored with {\\it FUSE}, lines of sight that are on average more reddened than lines of sight in previous {\\it Copernicus} studies. We present three major results. First, we observe the well-established correlation between iron depletion and $\\nHavg$ and also find trends between iron depletion and other line of sight parameters (e.g.~$\\fHmol$, $\\ebv$, and $\\av$), and examine the significance of these trends. Of note, a few of our lines of sight probe larger densities than previously explored and we do not see significantly enhanced depletion effects. Second, we present two detections of an extremely weak Fe II line at 1901.773 \\AA{} in the archival STIS spectra of two lines of sight (HD 24534 and HD 93222). We compare these detections to the column densities derived through {\\it FUSE} spectra and comment on the line's $f$-value and utility for future studies of Fe II. Lastly, we present strong anecdotal evidence that the Fe II $f$-values derived empirically in through {\\it FUSE} data are more accurate than previous values that have been theoretically calculated, with the probable exception of $f_{1112}$. ", "introduction": "\\label{s:FeII_intro} Iron is both a relatively abundant element in the Galaxy and a major constituent in most grain models. Fe I has an ionization potential of only 7.87 eV, while Fe II has an ionization potential of 16.18 eV. Therefore, the dominant form of gas-phase interstellar iron found in H I regions should be Fe II. In H II regions, iron should be found as a mix of Fe II, Fe III, and Fe IV. In $\\Hmol$ regions, some Fe I may exist. The average gas-phase abundance of iron that has been established implies that the vast majority of interstellar iron is tied up in dust. However, observed variations in the iron abundance, while having only marginal implications for the dominant ingredients of dust, may still shed light on the physical conditions of interstellar clouds. Three major studies of interstellar gas-phase iron have been carried out: \\citet{SavageBohlin} and \\citet*{Jenkins1986} with {\\it Copernicus} data; and \\citet*{SRF2002} with {\\it FUSE} data. (For the rest of this paper, these three papers will be referred to as SB1979, JSS1986, and SRF2002, respectively.) \\citet{Howk} also studied Fe II with {\\it FUSE} but for the purpose of empirically determining oscillator strengths ($f$-values) in the {\\it FUSE} wavelength region rather than studying abundances and depletions. In summarizing previous studies and discussing our own results, it will be useful to define a few terms. By abundance, we mean the ratio of an element relative to hydrogen, given either logarithmically or linearly. We will define the iron abundance as Fe II/$\\Htot=\\NFeII/[\\NHI + 2\\NHmol]$, and neglect other ionization states of iron and hydrogen for the reasons discussed above. The quantity $\\delta$ is the ratio of the gas-phase abundance seen in a particular line of sight to an assumed cosmic abundance standard. Despite difference in terminology in the literature (cf.~SB1979 and SRF2002), we will refer to the quantity $D=-\\log{\\delta}$ as the depletion. It is the logarithm of the ratio of the total amount of an element to the amount in the gas-phase. Therefore, an increase in depletion means an increase in the amount of that element tied up in dust grains and/or molecules. This definition of depletion is useful because errors in the assumed cosmic standard produce a constant shift in the quantity $D$ and trends can still be analyzed. SB1979 found the interstellar ratio of iron to hydrogen to be equal to $4.5\\times10^{-7}$. They observed depletions in the range of $D=0.11-2.49$. SB1979 found positive correlations between $\\NFeII$ and the quantities $\\NHtot$ and $\\ebv$. These results do not consider depletion, and could be consistent with constant depletion. However, they did find iron depletion to be correlated, to varying degrees, with several parameters. The first very rough correlation they cite is between depletion and $E_{1330-V}/\\ebv$, a measure of dust grain size. SB1979 suggested that this implies that small grains are removed through grain growth and coalescence as opposed to destruction. However, other dust-related parameters such as visual polarization from grains and the gas-to-dust ratio (as measured by $\\NHtot / \\ebv$) were not found to correlate with iron depletion. Clearer correlations did exist for $\\nHavg$ ($\\equiv \\NHtot/r$, where $r$ is the line-of-sight pathlength) and $\\ebv/r$, two quantities that should show similar correlations because of the correlation between $\\NHtot$ and $\\ebv$ \\citep{Bohlin1978}. In these correlations, iron depletion increases with increased average density in the line of sight. Two other measures of cloud conditions that SB1979 found to be correlated to iron depletion were the molecular fraction of hydrogen, $\\fHmol$---positively correlated with depletion---and kinetic temperatue, $T_K$---negatively correlated with depletion. These correlations were somewhat rough, however. All of these correlations ($\\nHavg$, $\\ebv/r$, $\\fHmol$, and $T_K$) imply that iron depletion increases in denser environments, possibly due to density-dependent formation or destruction processes or shielding from $\\Hmol$. Finally, SB1979 also found a correlation between the depletions of Fe and Ti, a result that, when combined with other results from the literature \\citep{deBoerLamers, Stokes}, suggests similarities between grains that contain Fe, Ca, Mn, and Ti. JSS1986 measured the abundances of several refractory elements (Mg II, P II, Cl I, Cl II, Mn II, Fe II, Cu II, and Ni II) with {\\it Copernicus}. The major conclusion of the JSS1986 study was that depletions of all the elements are strongly correlated with the average hydrogen density in the line of sight, $\\nHavg$, with increasing depletion related to increased density. Depletions of various elements are also correlated with each other. JSS1986 interpreted these results largely in the context of the model of \\citet{Spitzer1985}, where the ISM is considered to be comprised of two types of clouds, warm and cold. According to this model, individual lines of sight should exhibit properties that are a result of sampling both types of clouds, and the best parameter for characterizing whether warm or cold clouds are sampled is $\\nHavg$. JSS1986 found $\\logFeIIH=-6.46\\pm0.06$ for cold clouds and $\\logFeIIH=-5.84\\pm0.05$ for warm clouds. The observed random deviations from a smooth trend in the relationship with $\\nHavg$ were larger than those expected from just the errors in measuring the iron and hydrogen column densities; JSS1986 used this fact to estimate that the true dispersion in the iron abundance is between 0.03 and 0.11 dex. It is worth noting that JSS1986 expressed two observational concerns that may have affected their data: (1) the possibility of ``hidden,'' highly-saturated components that, though undetected, might contribute significantly to the measured column densities and (2) significant contributions to the column densities from H II regions that violate the assumption that the true iron abundance, $\\NFe/\\NHtot$, is accurately represented by $\\NFeII / [\\NHI + \\NHmol]$. SRF2002 examined Fe II abundances with {\\it FUSE}, and explored 18 reddened lines of sight with greater average extinction, reddening, and molecular fraction of hydrogen than the SB1979 and JSS1986 surveys, though the SRF2002 sample did not probe lines of sight with values of $\\nHavg$ that were significantly larger than the densest lines of sight in SB1979 and JSS1986. SRF2002 found that while Fe II depletion increases with increased $\\ebv$ up to about $\\ebv \\approx 0.35\\magnitude$, primarily in the {\\it Copernicus} targets, the trend levels off at larger values of $\\ebv$ (see Fig. 3 of SRF2002). No correlation was found between depletion and $\\av$ (Fig. 4 of SRF2002), for a wide range of extinction ($\\av \\approx 1-3.5\\magnitude$); in fact, the depletions appear to be very constant. SRF2002 similarly concluded that there is little correlation between iron depletion and the molecular fraction of hydrogen (Fig. 5 of SRF2002, including the results of SB1979). Lastly, Fig. 6 of SRF2002 shows the correlation of iron depletion with average line of sight density, $\\nHavg$. That figure shows that the SRF2002 results are consistent with the correlation already observed by SB1979 and JSS1986, though, as stated above, SRF2002 did not probe larger $\\nHavg$ than those surveys. SRF2002 discussed in detail potential interpretations of the observed correlations (or lack thereof) and the fact that significantly enhanced depletions are $not$ observed in clouds with large $\\ebv$, $\\av$, and $\\fHmol$, contrary to what was expected prior to the survey. That discussion involved the creation and destruction of grain mantles, the warm and cold cloud mixing model of the ISM suggested by \\citet{Spitzer1985}, and the definition of {\\it translucent lines of sight} as opposed to {\\it translucent clouds}. This latter idea is also discussed in \\citet{SnowMcCall}, with the principle being that diffuse material can make a significant contribution to the reddening and extinction in a line of sight that might otherwise be assumed to match the assumed characteristics of a ``translucent cloud.'' We will return to some of these ideas in \\S \\ref{s:FeII_results}. A more recent but smaller survey by \\citet{Miller2007} examined the iron and silicon abundances in six lines of sight characterized as ``translucent.'' \\citeauthor{Miller2007}~found that both iron and silicon depletion increases in lines of sight with increased values of the extinction parameter $c_4$ \\citep[in the scheme of][]{FM1988}. Larger values of $c_4$ should correspond to an increase in the population of small grains. Therefore, \\citet{Miller2007} suggest that when additional depletion of iron and silicon is observed, those elements have been incorporated into smaller grains. It is important to note that \\citet{Miller2007} used weak lines in the STIS wavelength region to measure the Fe II abundance. Four of the six lines of sight from \\citet{Miller2007} have also been observed with {\\it FUSE} and have measurements of the iron abundance found in either SRF2002 (HD 27778 and HD 207198) or this paper (HD 147888 and HD 152590). We have carried out this study for the following reasons (roughly in decreasing order of importance): \\begin{enumerate} \\item To increase the sample size of reddened lines of sight with measurements of iron abundances and depletions. All of our lines of sight have $\\logHtot \\gtrsim 21$. Additionally, $\\sim \\frac{2}{3}$ of our lines of sight are characterized by $\\av \\gtrsim 1$, $\\sim \\frac{2}{3}$ by $\\ebv \\gtrsim 0.3$, and $\\sim \\frac{2}{3}$ by $\\fHmol \\gtrsim 0.1$. In conjunction with the 16 unique lines of sight from SRF2002---we have reanalyzed HD 24534 and HD 73882 in this paper---our data nearly quadruples the size of the overall sample of reddened lines of sight examined with {\\it FUSE}. Additionally, while most of the average line of sight densities in our sample are not significantly denser than the densities in previous studies, two lines of sight, HD 147888 and HD 179406, do probe larger $\\nHavg$ than any line of sight in SRF2002. \\item To probe lines of sight with less extinction and/or reddening than SRF2002. This is useful because certain extinction parameters (specifically $\\av$ and $\\rv$) were not available in many cases for {\\it Copernicus} targets (cf.~Figure 4 of SRF2002). In addition, many {\\it Copernicus} targets were short-pathlength lines of sight toward bright stars; using {\\it FUSE}, we are able to probe fainter targets and longer pathlengths (though there is some overlap between the two samples). \\item To present two detections of a very weak line of Fe II ($\\lambda_{\\rm rest}$=1901.773 \\AA{}) in archival STIS spectra (lines of sight toward HD 24534 and HD 93222). We place limits on the $f$-value of this line and discuss its utility for future determinations of Fe II. \\item To evaluate the $f$-values of the Fe II used in this study, comparing the empirical values from \\citet{Howk} and the theoretical calculations of \\citet{RU1998}. \\item To make simple models of Fe II in several lines of sight where we are also studying the 1260 \\AA{} line (along with other absorption lines) of Si II (A. G. Jensen \\& T. P. Snow, in preparation). \\end{enumerate} In \\S \\ref{s:FeII_obsdata} we discuss our observations and data reduction, including comments on the absorption lines used and our measurement methods. In \\S \\ref{s:methods} we discuss our methods of deriving column densities, abundances, and depletions. In \\S \\ref{s:FeII_results} we discuss our results and in \\S \\ref{s:FeII_summary} we summarize. ", "conclusions": "\\label{s:FeII_results} As stated above, our derived column densities are given in Table \\ref{coltable}, while curves of growth can be seen in Figures \\ref{fig:cogs1-15}-\\ref{fig:cogs46-51}. For our sample (not including the unique lines of sight from SRF2002) we find that Fe II/$\\Htot=3.2\\pm0.1\\times10^{-7}$, an average weighted by the inverse squares of the errors in the abundances. This is somewhat smaller than the results of SB1979, Fe II/$\\Htot=4.5\\times10^{-7}$; however, the two numbers are calculated differently. Our calculation is a weighted average of the abundances in each line of sight; the SB1979 abundance is a ``total'' abundance, calculated from the ratio of total observed Fe II to total observed hydrogen. If we calculate a total abundance in this manner, we obtain a total abundance that is larger, Fe II/$\\Htot=5.6\\times10^{-7}$. The weighted average tends to favor the smaller abundances which also have smaller errors, whereas calculating a total abundance slightly favors the tail of larger abundances, which are up to a few times larger than the median. Whether we consider an average or the total iron abundance, our results are distinctly different from the results of SRF2002; the 16 lines of sight unique to that study have a weighted average of $1.9\\pm0.2\\times10^{-7}$ and a total abundance of $3.5\\times10^{-7}$. The fact that our abundances are larger in both cases has a simple interpretation---our study covers some of the same highly-reddened ground as SRF2002, but with an additional sample of much less-reddened lines of sight. We note here that the averages of JSS1986, discussed in \\S \\ref{s:FeII_intro}, correspond to $1.4^{+0.2}_{-0.1}\\times10^{-6}$ for the ``warm'' ISM and $3.5\\pm0.5\\times10^{-7}$ for the ``cold'' ISM; our weighted average is similar to the latter sample, but our total abundance falls squarely between the two (as it should if our sample covers both ranges of the two JSS1986 samples). Much more insightful than the averages, however, are that we also see several statistically significant correlations between iron depletion and various line of sight parameters; these correlations and their interpretations are discussed in \\S \\ref{ss:FeII_correlations}. In these plots and analyses, we include the results of SRF2002 but not SB1979 or JSS1986. As discussed in \\S \\ref{s:FeII_intro}, our sample, compared to these previous samples, is either larger or covers more parameter space. There are potential systematic errors between the {\\it FUSE} studies (this paper and SRF2002) and SB1979 and JSS1986. SB1979 used the 1122.0, 1133.7, and 1144.9 \\AA{} lines of Fe II, while JSS1986 used the 1096.9, 1122.0, and 1133.7 \\AA{} lines. In addition to the differences in absorption line sets between those studies and ours, SB1979 used $f_{1133.7}=6.3\\times10^{-3}$ and $f_{1144.9}=0.15$, while JSS1986 adopted $f_{1133.7}=4.8\\times10^{-3}$. As noted in Table \\ref{linetable}, we used the $f$-values adopted by \\citet{Howk} of $f_{1133.7}=5.5\\times10^{-3}$ and $f_{1144.9}=0.106$. However, in comparing their results with SB1979, SRF2002 concluded that the SB1979 abundances should be increased (due to the downward revision of $f$-values) by an average of 20\\%, much smaller than the inherent variation in abundances. Therefore, there is not a clear reason to suspect that $f$-value differences lead to significant systematic errors. On the other hand, the fact that this study and SRF2002 are based on more absorption lines (7 in most cases) covering a broader range in $f$-value than the studies by SB1979 and JSS1986 (3 lines at most) is one reason to prefer the {\\it FUSE} results; at the least, it justifies including only the SRF2002 results in our analysis. Our search for the weak 1901.773 \\AA{} line of Fe II and the two probable detections that resulted are discussed in \\S \\ref{ss:1902line}. A brief comparison of the previous theoretical \\citep{RU1998} and empirical \\citep{Howk} $f$-values is given in \\S \\ref{ss:f-values}. Finally, we discuss the overall Galactic abundance of iron (and, briefly, the implications for dust models) in \\S \\ref{ss:cosmic_iron}. \\subsection{Correlations} \\label{ss:FeII_correlations} In this subsection we discuss the correlations and anticorrelations found between iron abundances and depletions and various line of sight parameters (measures of hydrogen content, extinction, reddening, and line of sight pathlength). As discussed, above, we have included the lines of sight from SRF2002. We use the Fe II column densities determined in that paper, but in order to calculate abundances and depletions and analyze correlations we find our own values from the literature for the line of sight parameters just mentioned. This is the reason for any observed discrepancies between our plots and the plots in SRF2002. We have not included the data points from either SB1979 or JSS1986 for the reasons discussed above. Correlations between iron depletion and measures of hydrogen (total hydrogen column density, average hydrogen volume density, and the molecular fraction of hydrogen) are discussed in \\S \\ref{sss:FeII_hydro_corr}; correlations between iron depletion and extinction and reddening parameters are discussed in \\S \\ref{sss:FeII_ext_corr}; an anticorrelation between iron depletion and distance (i.e.~line of sight pathlength) is discussed in \\S \\ref{sss:FeII_dist_anticorr}; and lastly, we provide a few comments on outlying data points in \\S \\ref{sss:outliers}.% \\subsubsection{Correlations with Hydrogen} \\label{sss:FeII_hydro_corr} Iron depletion shows a clear correlation with increased average hydrogen volume density; see Figure \\ref{fig:logFeIIHlognh}. This is not surprising; this trend was already known for Fe II (JSS1986, SRF2002), and known or at least suggested for several other elements (JSS1986; \\citeauthor{Cartledge2001}~\\citeyear{Cartledge2001}, \\citeyear{Cartledge2004}). The physical explanation usually invoked for this phenomenon is the model by \\citet{Spitzer1985}, discussed in \\S \\ref{s:FeII_intro}, that measurements of individual lines of sight are probing two fundamental cloud types, warm and cold, and that the relative mix of the two cloud types along the line of sight determines the observed depletions. The average hydrogen volume density, $\\nHavg$, is usually taken to be the best diagnostic for the determining the mix of the two cloud types being probed. In general, we have not probed significantly larger $\\nHavg$ than previous studies, with one exception: HD 147888 has an average hydrogen volume density of 13.6 cm$^{-3}$, more than twice as dense than any of the lines of sight in this sample, SRF2002, or SB1979. (Note that HD 179406 is also more dense than any line of sight in SRF2002, though only half as dense as HD 147888.) While 147888 is among the most highly depleted lines of sight in this sample ($\\logFeIIH=-6.87$), it does not show depletion that is particularly extreme; seven other lines of sight have $\\logFeIIH<-6.8$ within their errors. Although we do not want to overstate the importance of one data point, this line of sight could potentially support the \\citet{Spitzer1985} model at least up to this larger $\\nHavg$. It would seem that even in this dense line of sight we are not probing clouds with the physical conditions expected of translucent clouds---or, perhaps, models of translucent clouds need to be adjusted to explain the observed lack of extreme depletions. We also find that iron depletion is correlated with the total hydrogen column density. \\citet{WakkerMathis} have suggested that $\\NHI$ is in and of itself a good indicator of depletions. While the correlations of depletion with respect to both $\\NHI$ and $\\nHavg$ are similar, we examined the partial correlation coefficient of the iron abundance and hydrogen volume density while hydrogen column density held constant (an example of using partial correlation coefficients in this manner is outlined in JSS1986). When we examine the correlation of the logarithmic iron abundance with the logarithm of $\\nHavg$ and $\\logHtot$, the partial correlation coefficient of the iron abundance and $\\nHavg$ is -0.447. Using a $t$-distribution for the appropriate number of degrees of freedom (in the manner of JSS1986), we find that the probability that the observed correlation coefficient has an absolute magnitude this large if the true correlation coefficient is 0 (we will refer to this as the probability of the null hypothesis) is 0.03\\%. If we examine the converse, the partial correlation between iron depletion and $\\logHtot$ with $\\lognHavg$ held fixed is still significant: the coefficient is -0.267; the two-sided probability of the null hypothesis is 3.4\\%. These results imply that while observationally speaking, both $\\lognHavg$ and $\\logHtot$ are very good predictors of abundances and depletions, $\\nHavg$ appears to have more fundamental physical significance. We note here that $\\logFeIIH$, $\\lognHavg$, and $\\logHtot$ were chosen as opposed to their linear counterparts because these variables show the strongest correlations. We also note that using these correlation coefficients makes several implicit assumptions, including a linear correlation between the variables being examined, and Gaussian distributions of the variables for the true overall populations. The validity of these assumptions potentially influences the validity of the above analysis, but we do conclude that $\\nHavg$ is more significant. SB1979 noted a possible rough correlation between iron depletion and the molecular fraction of hydrogen, $\\fHmol$. However, SRF2002, including the SB1979 points, did not conclude that there was an overall correlation. Examining Fig. 5 of SRF2002 reveals that in general, the lines of sight with a large molecular fraction (primarily the {\\it FUSE} lines of sight) have relatively constant depletion, while lines of sight with a smaller molecular fraction (primarily the {\\it Copernicus} lines of sight) show a great deal of scatter in iron depletion (over an order of magnitude in the iron abundance). Analyzing only our data points and those from SRF2002, we {\\it do} observe an overall correlation between iron depletion and $\\fHmol$; see Figure \\ref{fig:logFeIIHHf}. With the exception of a few lines of sight, this is in fact one of the strongest correlations seen in our data. We have briefly explored the possibility that in regions with a large $\\fHmol$ this additional iron is found in the form of gas-phase Fe I, but conclude that the gas-phase Fe I abundance is, in general, at least an order of magnitude too small to account for the difference. Despite the strength of the overall correlation, there are several outlying points. Lines of sight with a large $\\fHmol$ but little depletion include HD 210121 (from our sample) and HD 27778 and HD 62542 (from SRF2002), while lines of sight with a small $\\fHmol$ but larger depletions include HD 147888, HD 152236, and HD 164740. We will return to these outlying data points in \\S \\ref{sss:outliers}. Ultimately we conclude that while $\\fHmol$ is strongly correlated with iron depletion, it is not a perfect indicator of the physical conditions required for large depletions, and environments that are merely very dense without necessarily forming significant $\\Hmol$ can still show large depletions. As above, to examine the independent significance of the $\\fHmol$ correlation, we examined the partial correlation coefficients between iron depletion and $\\fHmol$ with $\\nHavg$ held fixed, we conclude that the correlation with $\\fHmol$ is largely secondary to the correlation with $\\nHavg$. If the three variables considered are $\\logFeIIH$, $\\fHmol$, and $\\lognHavg$, the probability of the null hypothesis is well over 50\\% whether $\\fHmol$ is considered linearly or logarithmically. If the linear values of the iron abundance or $\\nHavg$ are used, the probability of the null hypothesis $\\nHavg$ decreases (i.e.~that the correlation is more real is likely); however, as before, $\\logFeIIH$ and $\\lognHavg$ are examined because these two variables exhibit a much stronger correlation than the linear variables. Again, we note the implicit assumption of linear correlations between whatever variables (linear, logarithmic, or otherwise) are under examination. In the sense that the correlation with $\\fHmol$ is not necessarily independently significant, we have not shown anything new. However, we have shown that in our sample a stronger correlation exists than in previously studied samples (e.g.~SB1979, JSS1986, SRF2002). Why the SB1979 sample shows such a large spread in depletion at lower values of $\\fHmol$ that is not seen in our sample is unclear. We have discussed earlier the potential for systematic errors between {\\it Copernicus} and {\\it FUSE} studies, but even if those errors are larger than we estimate, they do not account for the spread within the self-consistent {\\it Copernicus} data. Scattered light, which was worse for {\\it Copernicus} than {\\it FUSE}, may be a small factor in causing some variations, but is almost certainly not responsible for the majority of the scatter. Inherent variation over the short pathlengths of some of the SB1979 targets may also be factor. It is worth noting that while our results show a few points of scatter at both large and small $\\fHmol$, the SB1979 scatter is exclusively at small $\\fHmol$. At $\\fHmol \\gtrsim 0.3-0.4$, there are no SB1979 targets with small depletions, but rather the few SB1979 targets with $\\fHmol$ this large more or less conform to the trend seen in our sample. This compares well with \\citet{Cartledge2006}, where trends of depletion for elements such as magnesium are also seen with respect to $\\fHmol$---though not as strong as between depletion and $\\nHavg$---and with some scatter at small $\\fHmol$. \\subsubsection{Correlations with Extinction and Reddening Parameters} \\label{sss:FeII_ext_corr} Figure \\ref{fig:logFeIIHred} shows $\\logFeIIH$ as a function of four different reddening or extinction parameters---the relative color excess between the $B$ and $V$ bands, $\\ebv$; $\\ebv$ scaled by line-of-sight pathlength, $\\ebvdist$ the total visual extinction, $\\av$; and $\\av$ scaled by line-of-sight pathlength, $\\avdist$. Both $\\ebv$ and $\\av$ are measured in magnitudes, and we calculate $\\ebvdist$ and $\\avdist$ in magnitudes per parsec. The parameter of $\\ebv$ is known to correlate well with $\\NHtot$, and is thought to be strongly correlated with the total dust column density in the line of sight. Given the correlation we see between iron depletion and $\\NHtot$, it is not surprising that we also see a correlation between iron depletion and $\\ebv$---denser environments are correlated with both iron depletion and total dust content. SRF2002 noted that the correlation between iron depletion and $\\ebv$ seems to break off at $\\ebv \\approx 0.35 \\magnitude$. We do not observe this trend. In our combined sample, we see a very strong correlation overall. The slope we see for the entire sample is about half as steep as the slope seen in SRF2002 for lines of sight with $\\ebv < 0.35$, but a few times steeper than the slope of the lines of sight with $\\ebv > 0.35$ in that paper. If we take the same approach as SRF2002 and look for the point where the correlation disappears for larger $\\ebv$, we must restrict the sample to the 23 lines of sight with $\\ebv \\geq 0.5\\magnitude$. In this sense, we still see the same effect noted in SRF2002, which they interpreted as a threshold effect, wherein the conditions that increase $\\ebv$ do not necessarily require additional iron per hydrogen atom. The parameter $\\av$ also correlates well with the total hydrogen column density and should also be correlated with the total dust column density. As discussed in \\S \\ref{s:FeII_intro}, SRF2002 did not find any correlation between iron depletion and $\\av$ for the 18 lines of sight in that paper. They were not able to compare their 18 lines of sight with the SB1979 or JSS1986 results because those papers did not report on $\\av$ for those lines of sight. However, taking our sample and the SRF2002 sample, we do find a correlation between iron depletion and $\\av$. The reason for this difference seems to be the wider range of $\\av$ examined by our sample. When we restrict the sample to lines of sight with $\\av \\geq 1.5\\magnitude$, the correlation weakens, and it disappears for lines of sight with $\\av \\geq 2\\magnitude$ (at this point, however, it should be noted that the restricted sample contains only 11 lines of sight). If $\\ebv$ and $\\av$ are correlated with the total dust column density, then dividing by the pathlength of the line of sight results in quantities ($\\ebvdist$ and $\\avdist$) that should be correlated with the total dust volume density. Not surprisingly, iron depletion is correlated with both of these quantities, at least as strongly as it is correlated with the corresponding extinction parameters integrated over the entire line of sight. Calculating partial correlation coefficients indicates that the probability of the null hypothesis for a correlation between iron depletion and $\\ebvdist$ with $\\nHavg$ held constant is between 10\\% and 20\\%, for all possible combinations of linear and logarithmic quantities. The relationship between the iron abundance and $\\avdist$ with $\\nHavg$ held constant is less clear; the probability of the null hypothesis is 42\\% if $\\avdist$ and $\\nHavg$ are logarithmic, but less than 1\\% if they are linear. Reversing the correlations, we see that iron depletion is always very well-correlated with $\\nHavg$ even when other variables are held constant. We therefore conclude that the most important variable of interest for determining iron depletion is still $\\nHavg$, but that these measures of dust density ($\\ebvdist$ and $\\avdist$) are also useful quantities that independently correlate with the iron depletion to a small degree. That the measures of dust density should correlate with iron depletion is also physically motivated in that the depleted material should be found in the dust. That $\\nHavg$ is still the variable most strongly correlated with iron depletion potentially indicates that the way iron is incorporated into the dust is nonuniform, and/or that the dust grains that contain some iron have variable extinction characteristics. Dividing these extinction parameters by line of sight pathlength also somewhat resolves the ``threshold'' effects seen for the integrated extinction parameters alone. Correlations, though both weaker and less significant, remain---rather than disappear---even when only the densest lines of sight are considered. However, the fact that the correlations do weaken and we do not see extreme depletions indicates that we are not observing lines of sight that are dominated by translucent clouds, though perhaps some lines of sight do probe these clouds to a limited extent. The quantity $\\rv$ is the ratio of total visual extinction to selective extinction ($\\rv$ is defined as $\\av / \\ebv$) and is correlated with grain size (because larger grains contribute significantly to $\\av$ but not to $\\ebv$). We find that while there is the hint of a correlation between iron depletion and with $\\rv$ (e.g.~a linear regression of the linear abundance and $\\rv$ has a statistically significant slope), upon further examination the correlation is not particularly significant. For example, using a Pearson correlation coefficient, the probability of the null hypothesis for logarithmic iron depletion and $\\rv$ is 35\\%. Rank correlation coefficients (Spearman's $\\rho$ and Kendall $\\tau$, using IDL's R\\_CORRELATE function) that do not carry any implicit assumptions about the functional form of the correlation also show that any correlation is slight at best. As discussed below in \\S \\ref{sss:outliers}, we also find anecdotal cases where lines of sight that do not conform to some of the observed trends (e.g.~$\\fHmol$) tend to be correlated with $\\rv$ values different from the interstellar average of 3.1; however, we cannot draw the conclusion that there is an overall correlation between depletion and $\\rv$. Lastly, we will point out a few interesting anecdotal cases regarding cases of extreme depletion and large reddening and extinction. Given that the overall trends hold, we could point out many cases, but we will restrict ourselves to the three cases with the largest depletions: from most to least depletion, HD 164740, HD 110432, and HD 147888. When considering the combined sample, HD 164740 has the largest $\\av$ and second largest $\\ebv$, HD 110432 has the second largest $\\avdist$, and HD 147888 has the largest $\\avdist$ and $\\ebvdist$. And though we do not claim a conclusive trend between depletion and $\\rv$ overall, we note that these three lines of sight all have large $\\rv$ of $\\gtrsim4$, including HD 164740 with the second largest $\\rv$ in the sample of 5.36. \\subsubsection{Anticorrelation with Distance} \\label{sss:FeII_dist_anticorr} We find that iron depletion is anticorrelated with distance, i.e.~line of sight pathlength. With the exception of a few outlying points (HD 210121 from our sample and HD 27778 and HD 62542 from SRF2002), iron depletion decreases out to line of sight pathlengths of about 2 kpc. For lines of sight with pathlengths from about 2--6 kpc, iron depletion appears constant to within $\\sim0.4\\dex$, significantly less variation than the decrease of an order of magnitude found in lines of sight with pathlengths less than 2 kpc (again, ignoring the three outlying points). The fact that we see constant depletion for long pathlengths is sensible for two reasons. First, lines of sight with pathlengths of at least a few kiloparsecs should be sampling the conditions of several clouds, and therefore the total iron abundances and depletions should remain relatively constant. Secondly, these longer lines of sight are, on average, less dense and less reddened per unit distance than the shorter lines of sight. However, it cannot be assumed that the lines of sight are truly uniform; slight variations, therefore, may nevertheless contribute to our observed correlations with other line of sight parameters (such as reddening and extinction). \\subsubsection{Outlying Data Points} \\label{sss:outliers} We have already mentioned six outlying data points. HD 27778, HD 62542, and HD 210121 have smaller iron depletions but larger values of $\\nHavg$, $\\fHmol$, $\\ebvdist$, and $\\avdist$, breaking the trends that we have just discussed. Conversely, HD 147888, HD 152236, and HD 164740 have large iron depletions despite small values of $\\fHmol$. We will first approach the somewhat easier task of interpreting the second set of outliers. HD 147888, HD 152236, and HD 164740 all have very large hydrogen column and/or volume densities. Additionally, these lines of sight all have relatively large values of extinction and reddening---HD 147888 has the largest $\\avdist$ and $\\ebvdist$ in the sample, while HD 152236 and HD 164740 both have large pathlength-integrated values of $\\av$ and $\\ebv$ (when scaled by pathlength, HD 152236 is in the upper half of the sample for both parameters and HD 164740 is in the upper third). Therefore, given the trends we see with respect to all of these parameters, it is not surprising that we see large iron depletions. To explain the low fractions of $\\Hmol$, we note that all three of these lines of sight have $\\rv$ larger, and in two cases much larger, than the interstellar average of 3.1: 4.06 for HD 147888, 3.29 for HD 152236, and 5.36 for HD 164740. Large values of $\\rv$ imply a large average grain size, but a large average grain size also implies that the $\\Hmol$ formation rate may be small because of the reduced surface area per unit volume \\citep[see][]{Snow1983}. Comparing the three lines of sight, HD 152236 has the smallest iron depletion, smallest $\\rv$, and largest $\\fHmol$, while HD 164740 has the greatest iron depletion, largest $\\rv$, and smallest $\\fHmol$, with HD 147888 in the middle in all cases. These trends fit our interpretation that these lines of sight have smaller molecular fractions of hydrogen due to increased grain size, but otherwise follow the trends between iron depletion and overall gas and dust density. Additionally, we should note that the increased grain size is unlikely to cause the additional iron depletion, because the reduced average grain size that suppresses $\\Hmol$ formation should also reduce the sticking rate of atoms and ions to grains. Rather, the increased grain size is likely a result of grain coagulation in environments where iron is already highly depleted. It is more difficult to interpret the other three lines of sight with relatively low iron depletions despite having large hydrogen volume densities, large molecular fractions of hydrogen, and larger values of reddening and extinction (both per unit length and pathlength integrated). We first note that all three of these lines of sight have relatively large errors in the iron abundance, so it is possible that these lines of sight are merely statistical fluctuations in the overall correlation. However, assuming this is not the case, there are a few comments that we can make about these lines of sight. The atomic hydrogen column density that we report for HD 210121 is based on a 21 cm measurement from \\citet{WF1992}. If this measurement is saturated, then both the iron abundance and the molecular fraction of hydrogen are somewhat smaller and can be partially reconciled with the correlation. We note that while the total hydrogen column density derived from adding the $\\Hmol$ measurement of \\citet{Rachford2002} and the $\\NHI$ measurement of \\citet{WF1992} implies $\\logHtot=21.00$, the relationship of \\citet{Bohlin1978}, which correlates $\\ebv$ with $\\Htot$, implies that $\\logHtot=21.37$. Also worth noting is that if the HD 210121 curve of growth is fit with the equivalent width measurements being artificially weighted equally (i.e.~an unweighted curve-of-growth fit), the result for the iron column density is substantially smaller (by $0.31\\dex$). Therefore, either an underestimated value of $\\NHtot$ or an overestimate iron column density could be responsible for an iron abundance that seemingly deviates from certain trends. It is worth noting that the difference between the weighted and unweighted fits is a phenomenon that is generally not observed in other lines of sight. We have used the Fe II column density for HD 27778 from SRF2002, but we note that \\citet{Miller2007} derived a much smaller Fe II column density (and resulting iron abundance) for this line of sight. \\citet{Miller2007} used the 2260 \\AA{} line to derive the Fe II column density in HD 27778; we do not have an adequate basis for comparing the use of this line (primarily in terms of its $f$-value) to our curve-of-growth results. However, we note than an adoption of the \\citet{Miller2007} column density would eliminate this line of sight from consideration as an ``outlier.'' It is worth noting that the abundances of molecules such C$_2$ and CN are relatively large in HD 27778 \\citep{Federman1994}, HD 62542 \\citep{Gredel1993}, and HD 210121 \\citep{Gredel1992}. The properties of the diffuse interstellar band (DIB) absorption features are somewhat unique in the HD 62542 \\citep{Snow2002, ABM2005} and HD 210121 \\citet{Thorburn2003} lines of sight, with unusually strong ``C$_2$ DIBs'' (DIBs correlated with C$_2$) and unusually weak ``classical'' DIBs (other DIBs not correlated with C$_2$). All three lines of sight also have $\\rv$ at least slightly less than the interstellar average of 3.1: 2.73 for HD 27778, 2.83 for HD 62542, and 2.07 for HD 210121. The last point, about the small values of $\\rv$, implies that the average grain size is smaller in these lines of sight. With a population of small grains, the surface area per unit volume is increased and the rate of $\\Hmol$ formation may increase. This may be the case for these lines of sight---supported by the fact that of the three lines of sight, HD 210121 has the smallest value of $\\rv$, and presumably the smallest average grain size, but the largest fraction of $\\Hmol$. Also potentially supporting this interpretation are the large abundances of other molecules in these lines of sight. However, an important caveat is that the increased surface area per unit volume should also increase sticking of gas-phase atoms to grains. Therefore, it is still unclear why the iron depletions are so low. If the smaller average grain size is the result of destruction processes (e.g.~shocks), then the answer may be that the destruction processes may somehow preferentially destroy iron-bearing grains, releasing iron back to the gas phase. Again, however, it is very unclear why this would be the case, as shocks are thought to destroy grain mantles rather than grain cores, and the major source of depletion for iron should be grain cores, not mantles. While it is beyond the scope of this paper to explore this issue in detail, we have also examined correlations between $\\rv$, $\\nHavg$, and $\\fHmol$ in our entire sample. We see, for instance, that $\\fHmol$ increases with $\\nHavg$ up to about $\\nHavg \\approx 2$ cm$^{-3}$, where the correlation begins to break down and, in fact, becomes an anticorrelation (though not as statistically significant as the correlation at low densities). Therefore, we are possibly seeing evidence of the low $\\Hmol$ formation rate in dense environments. There is a significant amount of scatter in our attempt to correlate $\\rv$ and $\\fHmol$, with most lines of sight centered near $\\rv=3.1$ with a wide range of $\\fHmol$. Nevertheless, there is an anticorrelation between $\\rv$ and $\\fHmol$, significant at the 1-$\\sigma$ level, that can be found using several different correlation methods---a simple linear regression, in addition to Pearson, Spearman's $\\rho$, and Kendall $\\tau$ correlation coefficients. This anticorrelation exists for both the entire range in $\\rv$ and for a narrower range in $\\rv$ near the ISM average of 3.1. Again, given that $\\rv$ is correlated with grain size, this could be interpreted as limited evidence that the $\\Hmol$ formation rate is lower in environments with larger grains. However, radiation may also be an important factor; this is briefly discussed below. While less relevant to the issue of $\\Hmol$ formation, we also note that there is not a statistically significant correlation between $\\nHavg$ and $\\rv$. However, in spite of the explanation that variations in the correlation between the iron abundance and $\\fHmol$ might be explained by unusual values of $\\rv$, we should also note that unusual values of $\\rv$ do not guarantee a break from this correlation. For example, HD 38087, with the largest $\\rv$ in our sample, does not break from the trend. This may be in part due to the fact that while $\\rv$ gives a rough measure of average grain size, it does not give precise information about the details a grain population, nor does it measure total grain surface area (as opposed to surface area per unit grain volume). For example, while grain coagulation reduces the surface area per unit volume, an increased average grain size due to mantle growth will increase the total grain surface area \\citep{Snow1983}. An additional consideration is $\\Hmol$-dissociating radiation. The HD 147888 line of sight passes through the $\\rho$ Oph cloud complex (HD 147888 is $\\rho$ Oph D), which is known to have a high internal ultraviolet radiation field. This may be generally true regarding lines of sight with large values of $\\rv$---for example, Figure 1 of \\citet{Draine2003}, calculated using the data of \\citet{Fitzpatrick1999}, shows that for the same value of $\\av$, a typical line of sight with $\\rv=4$ has significantly less extinction (about 3 magnitudes) at 1000 \\AA{} than a typical line of sight with $\\rv=3.1$. Thus, a higher far-UV radiation field allowed by the reduced far-UV extinction almost certainly plays some role in the small values of $\\Hmol$ in some lines of sight with large values of $\\rv$. In some cases, the far-UV radiation may be the dominant cause. However, it should be noted that \\citet{Snow1983} nevertheless argues, based on the large number of lines of sight that are both dense and have small values of $\\fHmol$ in the survey of \\citet{ChaffeeWhite}, that a high radiation field is unlikely to be exclusively responsible for these cases. In conclusion regarding the issue of grain size affecting $\\Hmol$ formation rates, we have shown a few interesting anecdotal cases wherein applying this hypothesis to these outlying points seems to have some explanatory power. However, this cannot necessarily be considered a uniform effect or the dominant cause of these outlying points. Making definitive predictions would require further information about the grain population (namely, total grain mass) and the local radiation field. \\subsection{The 1901.773 \\AA{} Line of Fe II} \\label{ss:1902line} As mentioned in \\S \\ref{ss:FeII_HSTdata}, Fe II has a very weak transition at 1901.773 \\AA{}. This wavelength region is within the STIS wavelength coverage for 17 lines of sight in the combined sample of this paper and SRF2002 (near the edge of coverage in many lines of sight). Prior to this paper, there have been no published detections of this line in interstellar absorption due to its small $f$-value \\citep[calculated to be $7.00\\times10^{-5}$ by][]{RU1998}, although weaker lines of Fe II at 2267 \\AA{} \\citep[$f=2.16\\times10^{-5}$;][]{CardelliSavage1995} and 2234 \\AA{} \\citep[$f=1.29\\times10^{-5}$;][]{Miller2007} have been detected. We examined the available STIS data (see \\S \\ref{ss:FeII_HSTdata} for our methods) in an attempt to detect this feature, and our search resulted in two detections (HD 24534 and HD 93222) and 15 upper limits. In this section, the detections will be discussed. Below (\\S \\ref{ss:f-values}) we will discuss the upper limits and the derivation of an $f$-value from these data. The results are recorded in Table \\ref{eqwidths1902}. In the spectrum of HD 24534, we find a small feature at 1901.85 \\AA{} with a central depth of $\\sim5\\%$ of the continuum that we identify as the 1901.773 \\AA{} line. We cite the following pieces of evidence in favor of this identification: (1) the velocity offset of $\\approx13\\kmpers$ precisely matches that of the dominant velocity component of other lines of Fe II and other elements seen in this line of sight, in particular the 1355.5977 \\AA{} O I and the 1608.4511 line of Fe II (the latter of which is too saturated and too similar in $f$-value to the 1144.9 \\AA{} line to be of additional use in this study); (2) the equivalent width is roughly consistent with the column density measurement, as discussed below; and (3) no other significant ground-state transitions of any abundant elements exist within 1 \\AA{} of the Fe II line. HD 24534 \\AA{} was studied in SRF2002, who found a Fe II column density of $\\logFeII=14.42^{+0.14}_{-0.13}$ and a $b$-value of $7.5^{+3.3}_{-2.0}\\kmpers$. Because we detected a feature that appears to be the 1901.773 \\AA{} line of Fe II, we have reanalyzed the {\\it FUSE} spectra to independently determine our own column density and $b$-value, $\\logFeII=14.63\\pm0.06$ and $b=6.4\\pm0.5\\kmpers$. Though the difference in the column density of $0.21\\dex$ is nontrivial, the column densities are nearly consistent within the 1-$\\sigma$ errors. The difference likely arises due to the fact that, because of our coadding procedures, we were able to detect the 1127 \\AA{} line of Fe II (the weakest of the {\\it FUSE} lines), which SRF2002 did not include in their analysis of HD 24534, and our equivalent widths are better constrained. Based on the revised column density that we have derived and the \\citet{RU1998} $f$-value of $7.00\\times10^{-5}$, the expected $\\eqw$ of the 1901.773 \\AA{} line for HD 24534 is 0.96 m\\AA{}. This is in excellent agreement with our measured equivalent width of $0.94\\pm0.17$ m\\AA{}. This fit, however, is subject to a possible systematic error. The area where we identify the feature is slightly asymmetric, and the fit of the line depends on whether we choose a narrow fit that focuses on the portion of the feature where the depth is the greatest or a broader fit that covers the entire feature, including the redward asymmetry. If we choose the latter fit, we obtain a larger $\\eqw$ of $1.73\\pm0.26$ (consistent with the other fit to within 2-$\\sigma$). Both fits have a similar reduced $\\chi^2$ of $\\approx 0.73$. We have selected the narrower fit because of its consistency with the column density, but note that this choice ignores this possible systematic error. We observe a similar feature in the spectrum of HD 93222 that we identify as this line. In this spectrum (with a lower S/N than HD 24534), the feature is at a central wavelength of 1901.62 \\AA{} and has a depth of $\\approx10\\%$ of the continuum. For the same three reasons as before (matching of velocity offset with other lines, rough consistency with derived column density, and lack of other potential identifications), we identify this feature as the 1901.773 \\AA{} line. The expected $\\eqw$ for HD 93222, based on the column density derived through the {\\it FUSE} lines, is 5.76 m\\AA{}. However, in HD 93222 we identify at least two marginally resolved cloud components in the FUV absorption lines; the velocity offset of the feature we detect matches the velocity offset of the weaker, narrower component. However, we have also analyzed the doublet of Mg II found at 1240 \\AA{}, and in that case find three major components---at -23, -6, and 7$\\kmpers$. The feature we detect and identify as the 1901.773 \\AA{} line is at -23$\\kmpers$ with respect to its rest wavelength; the -23$\\kmpers$ component is the strongest (i.e.~largest column density) component for Mg II. If we assume that the velocity structure of Mg II and Fe II is similar, then the stronger Fe II component (at $\\sim0\\kmpers$) in the FUV lines is not a single stronger component but a blend of the two weaker components seen for Mg II (-6 and 7$\\kmpers$), and the weaker Fe II component is identified with the strongest Mg II component (both at -23$\\kmpers$). We determined that the -23$\\kmpers$ component contains approximately 43\\% of the total column Mg II density. Assuming the same distribution for Fe II implies an expected equivalent width of 2.45 m\\AA{} for the -23$\\kmpers$ component, which agrees within the errors with our measured value of $1.95\\pm0.76$ m\\AA{}. Our simple simulations show that the smaller components (at -6 and 7$\\kmpers$) of the 1901.773 \\AA{} line may be simply lost in the noise. Both of these comparisons are based on the assumption that the \\citet{RU1998} $f$-value is correct. Working the above problem in reverse, we have taken our equivalent widths and column densities and calculated an empirical $f$-value in \\S \\ref{ss:f-values}. Although the value we calculate is consistent with the \\citet{RU1998} value, this line is worth more study to further constrain its $f$-value. In any case, these detections at the very least roughly confirm the line's $f$-value, which in turn is extremely important for future work with COS, now scheduled to be installed on {\\it HST} in 2008. COS will have greatly increased sensitivity (depending on wavelength) but somewhat lower resolution than the highest dispersion modes of STIS. It is possible that this line will be able to reveal Fe II column densities down to at least $\\logFeII\\approx14$ and possibly smaller, without fear of saturation effects. Detecting such small column densities with COS may not be necessary, however, as many COS targets will have much larger total hydrogen column densities; thus, iron column densities will also be larger unless there are enhanced depletion effects. If enhanced depletion effects are observed, however, this line may be of utility as our rough measure of its $f$-value is larger than the $f$-value of the weak Fe II lines used by \\citet{Miller2007}. Therefore, this line may be more easily detected in lines of sight with extreme depletion and/or poor S/N, while still being weak enough to guarantee reasonable freedom from saturation. \\subsection{Fe II $f$-values} \\label{ss:f-values} \\citet{Morton2003}, which is a definitive compilation of $f$-values and other atomic data for UV lines of astrophysical interest, quotes the $f$-values theoretically calculated by \\citet{RU1998} rather than the more recent empirical $f$-values derived by \\citet{Howk} using {\\it FUSE} data. However, SRF2002 used the \\citeauthor{Howk}~$f$-values in their study. Figures \\ref{fig:cogs1-15}-\\ref{fig:cogs46-51} show significant anecdotal evidence that the cases with the smallest errors in the equivalent widths produce extremely self-consistent curves of growth using the \\citet{Howk} $f$-values. The same holds true, in large part, for the curves of growth in SRF2002. We have carried through our curve-of-growth method using both the \\citet{Howk} and \\citet{RU1998}. In 39 of 51 cases, using the \\citet{RU1998} $f$-values produces values of $\\chi^2$ that are larger than the $\\chi^2$ values that result from using the \\citet{Howk} values. When comparing the 102 cases (51 lines of sight using each set of $f$-values), the 10 poorest fits use the \\citet{RU1998} $f$-values. It is also worth noting that in several cases the \\citet{RU1998} $f$-values produce a column density of Fe II that is unreasonably large---more than an order of magnitude larger than any of the column densities that are produced using the \\citet{Howk} $f$-values, and producing gas-phase abundances on the order of $10^{-5}$, much larger than has been historically observed. Though some of the initial calculations using the \\citet{Howk} $f$-values similarly produce very large column densities, an alternate solution (i.e., values in the $\\chi^2$ array that are separated in parameter space from the best solution, but where $\\chi^2$ is below the 1-$\\sigma$ cutoff) always presents itself. This is not true of the calculations using the \\citet{RU1998} $f$-values, where 15 of 51 lines of sight (including 4 of the 12 cases where the \\citeauthor{RU1998} $f$-values improve $\\chi^2$ compared to using the \\citeauthor{Howk} $f$-values) fail to have a reasonable solution. In another four of the 12 cases where use of the \\citet{RU1998} $f$-values improves the $\\chi^2$ compared to the \\citet{Howk} $f$-values the improvement is a factor of two or more. In these cases, however, the reduced $\\chi^2$ values using the \\citet{Howk} $f$-values are $\\lesssim1$, with a maximum of 1.37 for HD 147888. The other cases (HD 41117, HD 168941, and HD 179406) generally have much larger errors in the equivalent widths (contributing to the small values of $\\chi^2$) and less than the full complement of lines measured (e.g.,~only three lines are measured for HD 41117). Figure \\ref{fig:cogsfval} shows a sample comparison of the best-fit curves of growth for HD 195965 using the \\citet{Howk} $f$-values and the \\citet{RU1998} $f$-values. We have selected HD 195965 because the error in the column density using the \\citet{Howk} $f$-values is among the smallest in our sample ($^{+0.03}_{-0.02}\\dex$); in addition, this line of sight shows very little evidence of a multiple-component velocity structure. Examining Figures \\ref{fig:cogs1-15}-\\ref{fig:cogs46-51}, shows that there are many, many cases where the \\citet{Howk} $f$-values produce self-consistent curves of growth, particularly when the equivalent width errors are small. It is very unlikely that the curves of growth would be this consistent across a wide range of lines of sight if the \\citet{Howk} $f$-values were not roughly correct. It is beyond the scope of this paper to present a more formal analysis of the $f$-values than what has been presented above (though we further discuss one specific exception below). We should note that the \\citet{Howk} and \\citet{RU1998} $f$-values are consistent with each other within their errors. However, the difference is nontrivial for constructing a curve of growth, as can be seen in Figure \\ref{fig:cogsfval}. The discrepancy in the two sets of $f$-values for the seven absorption lines in this study ranges from $\\approx5\\%$ to a factor of 2.5; on average, the difference is 30\\%, or $0.11\\dex$. The consistency of our best curves of growth, however, argues that the \\citet{Howk} $f$-values are much more accurate than this, perhaps to within 0.02 or $0.03\\dex$ (5-7\\%). However, this is more of an argument for the self-consistency of the \\citet{Howk} $f$-values, rather than a strict constraint on the accuracy of their collective magnitude. A uniform error in the magnitude of the $f$-values would affect our derived column densities but would not significantly affect the relative correlations discussed in \\S \\ref{ss:FeII_correlations}. The main exception to our argument for the self-consistency of the \\citet{Howk} $f$-values is the 1112 \\AA{} line. Using these $f$-values implies that the 1112 \\AA{} line should have a larger $\\eqw$ than the 1133 \\AA{} line by 9\\% in the linear case (where $\\eqw \\propto f\\lambda^2$). In the 15 lines of sight where SRF2002 measured both lines, $W_{1133} > W_{1112}$ in 11 cases, though the 1-$\\sigma$ errors in $\\eqw$ do overlap in 14 of the 15 cases ($W_{1133} > W_{1112}$ for HD 197512 even when errors are taken into account). However, the weighted averages tell a different story: $W_{1133}=33\\pm1$ m\\AA{}, while $W_{1112}=27\\pm1$ m\\AA{}. We find essentially the same result in our data. Figures \\ref{fig:cogs1-15}-\\ref{fig:cogs46-51} show that in the cases with the smallest errors in the equivalent widths and the most self-consistent curve of growth, the 1112 \\AA{} line nearly always deviates from the curve of growth. Our weighted averages are $W_{1133}=49\\pm1$ m\\AA{} and $W_{1112}=46\\pm1$ m\\AA{} (the smaller difference between the two in our sample may be due to the fact that with increased $\\eqw$, the 1133 \\AA{} line is slightly saturated in more cases). Interestingly, \\citet{RU1998} did find that $f_{1112} < f_{1133}$. Again, we chose HD 195965 to examine this effect. If we fit a curve of growth while omitting the 1112 \\AA{} line, we find a slightly revised column density ($\\logFeII=14.87\\pm0.03$ as opposed to $\\logFeII=14.85^{+0.03}_{-0.02}$). If we then fit the 1112 \\AA{} line to this curve, we find that $f_{1112}=4.61\\times10^{-3}$, in much better agreement with \\citeauthor{RU1998} than \\citeauthor{Howk}~(see Table \\ref{linetable}). In test cases, revising the $f$-value of the 1112 \\AA{} line does not significantly alter our curve-of-growth results; therefore, we have not reanalyzed our column densities in light of this potential $f$-value revision. We are also presenting the first published detections in interstellar absorption of the 1901.773 \\AA{} line of Fe II. Therefore, we can use our results to constrain the $f$-value of this line. If $\\eqw$ and $\\lambda$ are in \\AA{} and $N$ is in cm$^{-2}$, then the following equation applies: \\begin{equation}\\label{eq:weakeqw} \\eqw=8.85\\times10^{-21}Nf\\lambda^2 \\end{equation} Solving for the $f$-value and substituting in the wavelength of the 1901.773 \\AA{} line, we find that $f=\\eqw/[3.20\\times10^{-14}N]$. Considering both the errors in $\\eqw$ and $\\logFeII$, our derived $f$-value is $6.9\\pm1.6\\times10^{-5}$ for HD 24534 and $5.5\\pm2.4\\times10^{-5}$ for HD 93222. The weighted average of these results is $6.5\\pm1.3\\times10^{-5}$. This agrees, within the errors, with the \\citet{RU1998} theoretical value of $7.00\\times10^{-5}$. Note, however, the assumptions that went into both equivalent width measurements (the choice between fits for the line in the HD 24534 spectrum, and the distinct cloud components assumed for HD 93222). At the suggestion of the anonymous referee, we have also taken a different approach to calculate the $f$-value of the 1901.773 \\AA{} line. In this approach, we attempt to measure $\\eqw$ for the lines even when no line is visibly apparent and there is no feature that will be fit by a program utilizing $\\chi^2$ minimization. This is done by using other lines as a proxy---the 1239.9523 \\AA{} line of Mg II \\citep[from our work in][]{JensenMgII} in most cases, and the 1608.4511 \\AA{} line of Fe II for 24534. The fits to these lines are examined for the range over which the profile model is less than 99\\% of the continuum. Given the saturation level in these lines, this cutoff value should include nearly all of the equivalent width. This wavelength range is then translated to the same velocity range near the 1901.773 \\AA{} line, and a measurement of $\\eqw$ is made by summing the depth of the data points relative to the assumed continuum. Fluctuations from the noise are assumed to cancel out to zero, and residual equivalent widths can be potentially observed. To calculate an error on $\\eqw$, we adapt the formula presented by \\citet{Jenkins1973} for the maximum equivalent width of an undetected absorption line: \\begin{equation}\\label{eq:upperlimits} W_{\\lambda,max}=\\frac{N_{\\sigma} d\\lambda \\sqrt{M}}{\\rm S/N} \\end{equation} In this equation, $W_{\\lambda,max}$ is the upper limit on $\\eqw$, $N_{\\sigma}$ is the number of $\\sigma$ confidence desired, $d\\lambda$ is the wavelength spacing of the pixels, $M$ is the number of consecutive pixels required for detection, and S/N is the signal-to-noise of the local continuum. We use translate this upper limit on the equivalent width into an error on the equivalent widths that we measure. In other words, we assume $W_{\\lambda,max}=\\sigma(W_{\\lambda})$ where the $W_{\\lambda,max}$ term is evaluated with $N_{\\sigma}=1$. Even though the resulting values of $\\eqw$ do not correspond to visually apparent features and are not statistically significant in many cases, the goal of this procedure is to use these measurements to derive an $f$-value for the line. The $f$-value is calculated in each case using Equation \\ref{eq:weakeqw}. The error in the $f$-value is carried through using standard error propagation, considering errors in both $\\eqw$ and column density. The number of pixels used in determining the error from Equation \\ref{eq:upperlimits} is the number of pixels that meet the condition mentioned above, that the range where the model fit is less than 99\\% of the continuum. Within this alternate method, we approach the problem in two different ways. The first way is to perform the summation over the entire range where all components of the proxy fit are less than 99\\%, calculating the $f$-value based on the column density derived through the {\\it FUSE} lines. In this case, the weighted average is $3.1^{+1.2}_{-1.1}\\times10^{-5}$. The other way is to perform the summation over only the velocity range where the dominant component is less than 99\\% of the continuum. We then derive the $f$-value using an adjusted column density, using the Fe II column densities of this paper (or SRF2002), but scaled by the relative fraction of the column density found in the dominant component of the proxy fit. Using this method, the weighted average of the fits is $6.5\\pm1.5\\times10^{-5}$. The fact that summing over the larger range results in a smaller weighted average shows that either (1) the noise fluctuations over such small ranges do not cancel on average or (2) there is some error in fitting the continuum. We prefer the method of directly analyzing the lines that are actually observed and fit with statistical significance. However, the alternate method of summing over the data points is also in rough agreement with the previously calculated value of the lines $f$-value, particularly when we only analyze the dominant component. Nevertheless, this method depends on calculating a statistically significant average from statistically insignificant individual measurements, and is subject to systematic error of making an assumption about what range over which the summation should be performed. In either case, as stated in \\S \\ref{ss:1902line}, we have made an important step in roughly confirming the $f$-value of this line. However, further study to improve these constraints is important so that this line can be used in future studies with COS. \\subsection{The Cosmic Value of Fe/H} \\label{ss:cosmic_iron} In the past several years, many papers have attempted to summarize the current knowledge of stellar and solar abundances, with implications for the cosmic abundance ``standards'' in the ISM, if consistent standards in fact exist. Three papers of particular importance for this discussion are \\citet{SnowWitt}, \\citet{SofiaMeyer}, and \\citet{Lodders}. \\citet{SnowWitt} argued for a cosmic iron abundance (including all iron in both gas and dust) of $\\logFeH=-4.57$, based on an average of field B stars ($\\logFeH=-4.51$), cluster B stars ($\\logFeH=-4.49$), and disk F and G stars ($\\logFeH=-4.74$). For iron (along with many other elements), \\citeauthor{SnowWitt} concluded that the Sun was substantially overabundant ($\\logFeH=-4.33$), an anomalous data point relative to the true Galactic abundances. \\citeauthor{SofiaMeyer}, examined B stars (without differentiation between cluster or field stars) and disk F and G stars, and found a weighted average of $\\logFeH=-4.55$ for each sample. The B star abundances of both \\citeauthor{SnowWitt} and \\citeauthor{SofiaMeyer} are reasonably consistent, but with a significant difference between the F and G star abundances; \\citeauthor{SofiaMeyer} explain this difference as a result of restricting their sample to stars with ages $\\leq 2$ Gyr. In any case, the final adopted cosmic iron abundance of both papers is consistent to with 0.02$\\dex$, well within the errors. Contrary to \\citeauthor{SnowWitt}, however, \\citet{SofiaMeyer} argued that the solar abundances may in fact reasonably represent cosmic abundances for many elements, in part due to downward revisions to solar abundances \\citep[mainly][]{Holweger}. \\citet{Lodders} summarized CI chondritic abundances and solar photospheric abundances, and used both to derive protosolar abundances. \\citeauthor{Lodders} adopts a protosolar iron abundance of $\\logFeH=-4.46$. The overall agreement (0.10$\\dex$---25\\%---not including the errors) between all types of measurements (stellar and solar) is the among the best for any of the elements discussed in \\citeauthor{SnowWitt} and \\citeauthor{SofiaMeyer}. Given this agreement, we can infer the amount of iron in dust fairly confidently. We adopt a cosmic abundance of iron of $3.1\\pm^{+2}_{-1}\\times10^{-5}$ (a weighted average of both \\citeauthor{SofiaMeyer} measurements and the \\citeauthor{Holweger} solar abundance). Our weighted average implies that only $\\sim$1\\% of the cosmic abundance in the gas-phase, and our maximum gas-phase abundances are at most approximately $2\\times10^{-6}$. This places a very tight constraint on iron that can be used in dust models \\citep*[e.g.][]{ZDA2004}. However, there are reasons to question any cosmic abundance standard, as we have discussed previously \\citep{Snow2000, JensenOI, JensenNI}---many processes, particularly those in stellar formation, could cause abundances in stars to deviate from the abundances of the ISM. SB1979 found some significant spatial variations in the Fe/H abundance---namely that Scorpius-Ophiuchus lines of sight were substantially more depleted than two lines of sight in Cygnus. In our sample, we do not note any particularly strong spatial effects. Most of our lines of sight are in or very near the disk ($|b|<10^{\\circ}$), and within that range there is no hint of a correlation between iron depletion and Galactic $b$. There are slight trends with Galactic longitude $l$ and overall location; in particular, many of the stars in the Crux/Musca region are somewhat less depleted ($\\logFeH \\approx -6$) than the average of our sample, though this is not true of the more reddened (and nearby) HD 110432 in the SRF2002 sample. However, this deviation is not substantial compared to the overall scatter in our sample. Iron abundances and depletions were explored in the past by SB1979 and JSS1986 with {\\it Copernicus} and more recently by SRF2002 with {\\it FUSE}. We have undertaken a survey that covers a wider range of reddening and extinction than covered in any of these studies. We find evidence that iron depletion correlates with many line of sight parameters (such as $\\nHavg$, $\\ebv$, $\\ebvdist$, $\\av$, $\\avdist$, and $\\fHmol$). Some of these correlations have been noted previously while others have not. The fact that we observe trends that have not been previously observed may be explained by the following: (1) this is the largest survey of interstellar Fe II yet performed with {\\it FUSE} and, in conjunction with data points from SRF2002, probes the widest range in line of sight parameters in a self-consistent manner; (2) {\\it Copernicus} suffered from scattered light to a greater degree than {\\it FUSE} (of order 10\\% for {\\it Copernicus} compared to only a few percent for {\\it FUSE}), which may be responsible, to a limited degree, for some scatter in the measured iron abundances; (3) many {\\it Copernicus} lines of sight had very short pathlengths, over which intrinsic scatter may be significant; and (4) quite simply, the {\\it Copernicus} lines of sight have not been analyzed in the light of some of these parameters, especially $\\av$. Correlations between iron depletion and $\\ebv$, $\\ebvdist$, $\\av$, $\\avdist$, and $\\fHmol$ can all be interpreted as related to the correlation between iron depletion and the average line of sight density $\\nHavg$. While a few of our lines of sight probe slightly larger density and/or extinction than previous studies, we do not see depletions that are particularly extreme relative to previously observed depletions, suggesting that at best our lines of sight only partially probe true translucent clouds, and are instead dominated by lines of sight with integrated ``translucent'' levels of reddening and extinction. Also of note are two detections of the 1901.773 \\AA{} line of Fe II in the spectra of HD 24534 and HD 93222. This very weak line is potentially very important for determining iron abundances in a straightforward fashion, without fear of saturation, with the Cosmic Origins Spectrograph. Further detections are needed to better constrain the $f$-value, but our detections make clear that the previously calculated theoretical value \\citep{RU1998} is at least roughly correct, and the line should be detectable at typical Fe II column densities, even in highly-reddened lines of sight with potentially extreme depletions, with the high S/N expected for COS. Finally, we briefly discuss the $f$-values of the FUV absorption lines of Fe II used in this study. Though the theoretical $f$-values of \\citet{RU1998} and the empirical $f$-values of \\citet{Howk} are consistent within the errors, the differences are nontrivial for constructing a curve of growth. We find that the \\citet{Howk} empirical $f$-values produce self-consistent curves of growth in far more cases than the \\citet{RU1998} $f$-values. We find an exception in the case of the 1112 \\AA{} line--to obtain self-consistent curves of growth; an $f$-value closer to the \\citet{RU1998} $f$-value is preferred." }, "0710/0710.3067_arXiv.txt": { "abstract": "The Parity doublet model containing the SU(2) multiplets including the baryons identified as the chiral partners of the nucleons is applied for neutron star matter. The chiral restoration is analyzed and the maximum mass of the star is calculated. ", "introduction": "The doublet parity model assumes, besides the nucleons, the presence of particles with opposite chirality named chiral partners. At high density environment as the one in neutron stars interior, there is the possibility of creating such heavy particles. The main difference between this model and the usual chiral model (Ref.~\\refcite{chi}) is that in this one there is a term of bare mass in the lagrangian density. This mass, called $M_O$ appears with fields that are a mixture of the fields of the particles and their chiral partners in such a way that it does not break chirality (Ref.\\refcite{par}). It is assumed that the star is in chemical equilibrium and the baryons interact through the mesons $\\sigma$, $\\omega$ and $\\rho$ (included in order to reproduce the high asymmetry between neutrons and protons). Electrons are included to insure charge neutrality. The lagrangian density of the system contains besides the kinetic and the interaction terms an explicit symmetry breaking term in order to reproduce the masses of the pseudo-scalar mesons. The coupling constants of the baryons are adjusted to reproduce the baryonic vacuum masses. In the high-density limit the nucleon and its chiral partner have degenerate masses ($M_0=790 MeV$) as the sigma field goes to zero and chiral symmetry is restored: \\begin{eqnarray} M^*_\\pm=\\sqrt{\\left[\\frac{(M_{N_+}+M_{N_-})^2}{4}-M_0^2\\right]\\frac{\\sigma^2}{\\sigma_0^2}+M_0^2}\\pm\\frac{M_{N_+}-M_{N_-}}{2}\\frac{\\sigma}{\\sigma_0}. \\end{eqnarray} A good candidate for the nucleon chiral parter is the N'(1535), but since the identification of chiral partners is still on its first steps we study the case with N'(1200) and N'(1500) for comparison. This variation has drastic consequences on the results. Four different cases first studdied in Ref.~\\refcite{par} are applied to neutron stars. ", "conclusions": "With increasing density .i.e. toward the center of the star, chiral partners begin to appear, reaching a point where they exist at the same rate as they corresponding particles. The decrease in the scalar condensates signals the restoration of the chiral phase. Depending on the parameters, the phase transition turns out to be a continuous cross-over or of first order for P4. The maximum mass of the star is higher when the coupling constant $g_4$ is set to zero (P1 and P3) so in order to reproduce the most massive star observed, that has $M=2.1^{+0.4}_{-0.5}M_{\\odot}$ (Ref.~\\refcite{21}), the best option would be the P1 description, because although the P3 predicts a higher maximum mass for the star, its phase transition occurs at a chemical potential bigger than 1700 MeV, too hight compared to predictions." }, "0710/0710.0702_arXiv.txt": { "abstract": "We present constraints on violations of Lorentz Invariance based on Lunar Laser Ranging (LLR) data. LLR measures the Earth-Moon separation by timing the round-trip travel of light between the two bodies, and is currently accurate to a few centimeters (parts in $10^{11}$ of the total distance). By analyzing archival LLR data under the Standard-Model Extension (SME) framework, we derived six observational constraints on dimensionless SME parameters that describe potential Lorentz-violation. We found no evidence for Lorentz violation at the $10^{-6}$ to $10^{-11}$ level in these parameters. ", "introduction": " ", "conclusions": "" }, "0710/0710.1018_arXiv.txt": { "abstract": "We show that near-infrared observations of the red side of the Ly$\\alpha$ line from a single gamma ray burst (GRB) afterglow cannot be used to constrain the global neutral fraction of the intergalactic medium (IGM), $\\bar{x}_H$, at the GRB's redshift to better than $\\delta \\bar{x}_H \\sim 0.3$. Some GRB sight-lines will encounter more neutral hydrogen than others at fixed $\\bar{x}_H$ owing to the patchiness of reionisation. GRBs during the epoch of reionisation will often bear no discernible signature of a neutral IGM in their afterglow spectra. We discuss the constraints on $\\bar{x}_H$ from the $z = 6.3$ burst, GRB050904, and quantify the probability of detecting a neutral IGM using future spectroscopic observations of high-redshift, near-infrared GRB afterglows. Assuming an observation with signal-to-noise similar to the Subaru FOCAS spectrum of GRB050904 and that the column density distribution of damped Ly$\\alpha$ absorbers is the same as measured at lower redshifts, a GRB from an epoch when $\\bar{x}_H = 0.5$ can be used to detect a partly neutral IGM at $97\\%$ confidence level $\\approx 10$\\% of the time (and, for an observation with three times the sensitivity, $\\approx 30$\\% of the time). ", "introduction": "Tomorrow, a gamma ray burst (GRB) may be observed that originates from the death of one of the first stars, during the epoch of reionisation. Despite the great distance to this burst, it will be the brightest gamma ray source on the sky for several tens of seconds, one of the brightest cosmological X-ray sources for hours, and its afterglow will be observable for weeks in the near-infrared (and, for the first few hours, brighter than any $z \\sim 6$ QSO). Much of the optical and near-infrared light will be obscured by the Ly$\\alpha$ forest, and this obscuration will enable the strongest constraint to date on the neutral hydrogen fraction of the intergalactic medium (IGM) at the burst's redshift. In fact, such an occurrence may already have been realised. \\citet{haislip06}, \\citet{kawai05}, \\citet{tagliaferri05}, and \\citet{totani06} observed and analysed the optical/near-infrared afterglow of GRB050904, identified to be at $z =6.3$ -- possibly during the reionisation epoch and the GRB with the highest confirmed redshift. \\citet{totani06} derived the constraint on the global neutral fraction $\\bar{x}_{H} < 0.6$ at $z = 6.3$. In this paper, we discuss the assumptions that went into their analysis, and we investigate how realistic modelling of reionisation can affect constraints on $\\bar{x}_H$ from GRB050904 and from future $z>6$ GRBs. The Swift satellite has greatly increased the sample of GRBs with known redshifts in the last two years \\citep{gehrels04}. Future missions such as EXIST \\citep{grindlay06} and JWST \\citep{gardner06} will further enhance our ability to detect high-redshift GRBs and will enable more detailed follow-up studies of their near-infrared afterglows. Interestingly, approximately one-half of Swift bursts are ``dark bursts'' -- bursts that have detected X-ray afterglows, but that have no measurable optical emission (e.g., \\citealt{filliatre06}). While it is probable that most dark bursts originate from low-redshift, dust-rich galaxies, a fraction of dark bursts may originate from $z>6$ and are ``dark'' because Ly$\\alpha$ absorption from the high-redshift IGM absorbs the optical emission (e.g., \\citealt{malesani05}). In addition to their extreme luminosity, there are several other advantages to studying reionisation with GRBs compared to other probes of this epoch. First, the afterglows of high-redshift GRBs are observed at earlier (brighter) times in the source frame than those at lower redshifts, so the dimming owing to increased luminosity distance is nearly cancelled, and the observed flux is almost independent of redshift \\citep{lamb01, ciardi99}. Second, unlike the spectra of galaxies and quasars, the intrinsic afterglow spectrum of a GRB is a featureless power-law at the relevant wavelengths, allowing a more precise measurement of absorption owing to a neutral IGM \\citep{barkana04b}. Finally, since the theoretical expectation is that most of the star formation at $z \\gtrsim 6$ occurs in halos with $m \\sim 10^9 \\; \\Msun$ and because observations at $z \\gtrsim 6$ currently probe only the most massive galaxies and QSOs ($m \\gtrsim 10^{11} ~\\Msun$), high-redshift GRB host galaxies should be less massive than galaxies selected in another manner. Consequently, GRB host galaxies will sit in smaller HII regions during reionisation (on average) than galaxies selected by different means. Therefore, GRBs will suffer a larger Ly$\\alpha$ IGM absorption feature. In this work, we do not concentrate on wavelengths blueward of source-frame Ly$\\alpha$ (in the Ly$\\alpha$ forest) to derive constraints from GRBs. Any blueward flux indicates the presence of ionised gas at the redshift of the transmission. However, at high redshifts there is little or no flux in the Ly$\\alpha$ forest, even in ionised regions, owing to the increase in density with increasing redshift, the decrease in the size and in the number of voids, and the decrease in the amplitude of the ionising background (e.g., \\citealt{becker06} and \\citealt{lidz07}). As a result, it is difficult to distinguish a partly ionised IGM from a fully ionised one with the $z > 6$ forest \\citep{becker06, lidz07}. Future observations of the $z >6$ Ly$\\alpha$ forest from additional QSOs and GRBs will aid reionisation studies, but is unclear whether such studies will ever provide definitive evidence for neutral pockets in the IGM. In contrast, the shape of the line profile redward of Ly$\\alpha$ \\emph{is} sensitive to a substantially neutral IGM and, therefore, can be used to unambiguously detect reionisation \\citep{miralda98}. Little is known about the rate of GRBs at $z > 6$. We assume that the rate of long GRBs traces the massive star formation rate (SFR) for most calculations in this work.\\footnote{We do not consider the other class of GRBs, the ``short'' GRBs, in this study. These bursts, while still cosmological, are more local than long GRBs and typically do not have a detected afterglow.} The assumption that the GRB rate traces the massive SFR is supported by observations of lower redshift GRB host galaxies \\citep{bloom02, djorgovski01}. However, \\citet{kistler07} found that the GRB rate is four times higher at $z \\approx 4$ than if the GRB rate exactly traces the SFR. Other properties of a galaxy apart from its massive SFR might be correlated with its rate of GRBs. For example, \\citet{stanek06} found that $z < 0.25$ GRBs -- GRBs that are typically under-luminous -- are preferentially in metal-poor galaxies. Making the assumption that the GRB rate traces the observed SFR, \\citet{salvaterra08} predicted that SWIFT will be triggered by $1-4$ bursts a year above $z = 6$\\footnote{This rate is larger than the SWIFT rate of $1$ identified $z >6$ bursts in three years, but it is possible that these numbers can be reconciled in light of these dark bursts.} and that the EXIST mission would observe $10-60$ bursts a year. Other studies have predicted even larger rates \\citep{bromm02, daigne06}. At $z > 6$, POPIII stars with average masses of $\\sim 100~\\Msun$ may exist. It is unclear whether the death of a POPIII star can result in a GRB. \\citet{fryer01} identified a mechanism that might produce GRBs from POPIII stars. However, once the interstellar metallicity reaches a critical value of $\\sim 10^{-3.5}$ solar in a high-redshift galaxy, POPIII star formation quenches and the normal mode of POPII star formation begins (e.g., \\citealt{mackey03, yoshida04}), and this mode {\\it is} known to produce GRBs. Most if not all of reionisation likely owes to photons from POPII-like stars (e.g., \\citealt{sokasian04, trac06}). In Section \\ref{redwing}, we discuss the absorption profiles of a neutral IGM as well as of a damped Ly$\\alpha$ absorber (DLA), and, in Section \\ref{reionisation}, we discuss our simulations of reionisation and their implications for the amount of IGM absorption in a GRB afterglow spectrum. Section \\ref{fits} quantifies the detectability of a neutral IGM from GRB afterglow spectra, and Section \\ref{GRB050904} describes the constraints on $\\bar{x}_H$ from the $z = 6.3$ burst, GRB050904. For our calculations, we adopt a cosmology with $\\Omega_m = 0.27$, $\\Omega_\\Lambda = 0.73$, $\\Omega_b = 0.46$, $\\sigma_8 = 0.8$, $n = 1$, and $h = 0.7$, which is consistent with the most recent cosmic microwave background and large scale structure data \\citep{spergel06}. We express all distances in comoving units unless otherwise noted. When this project was nearing completion, we learned of a similar effort by \\citet{mesinger07b} and refer the reader there for a complementary discussion. ", "conclusions": "GRBs are the most luminous sources at high redshifts, and their smooth power-law spectra are ideal for isolating the effects of absorption owing to a neutral IGM. However, observations must separate the impact of IGM absorption from that of a DLA to detect a neutral IGM. If no damping wing feature from IGM absorption is detected in the spectrum of a high-redshift GRB, one must be careful to conclude that reionisation is complete at the redshift of interest. We have shown that there is a wide probability distribution of HII region sizes, and that there is large variation in the distribution of $\\bar{x}_H$ along different sight-lines. A non-detection of neutral hydrogen from a GRB afterglow spectrum might arise because the GRB host galaxy sits within a large HII region. If absorption owing to a neutral IGM is detected, it will be impossible to infer $\\bar{x}_H$ from a single GRB to better than $\\delta \\bar{x}_H \\sim 0.3$ because of the patchiness of reionisation. Assuming an observation with similar sensitivity to the Subaru FOCAS spectrum of GRB050904, that the distribution of DLAs is the same as found at lower redshifts, and that the redshift of the GRB is known, a GRB from a redshift at which $\\bar{x}_H \\approx 0.5$ can be used to detect a partly neutral IGM at $98\\%$ C.L. $\\approx 10\\%$ of the time (and, for an observation with $3$ times the sensitivity, $\\approx 30\\%$ of the time). If $\\bar{x}_H < 0.5$, these percentiles for detection are even smaller. Weaker DLAs enhance the probability of detecting a neutral IGM, but too weak of a DLA may prevent a precise redshift determination, which is essential for tight constraints on $\\bar{x}_H$. Since the $z = 6.3$ burst GRB050904 has a DLA with $N_{\\rm HI} = 10^{21.6}$ cm$^{-2}$, the absorption on the red side of the line is dominated by the DLA \\citep{totani06}. While this burst may favour a model with an ionised universe over a neutral universe \\citep{totani06}, a weaker DLA is necessary to be able to constrain $\\bar{x}_H$ in the spectrum of a high-redshift GRB. This is particularly true if the GRB occurs within a large HII region. GRB050904 was observed spectroscopically by the Subaru telescope $3.4$ days after the prompt gamma ray emission. If this afterglow had been observed hours after the burst, the flux would have been more than an order of magnitude larger. To detect a neutral IGM, it is crucial for optical and near-infrared spectrographs to observe candidate high-redshift GRBs as soon as possible after the prompt gamma ray emission. A high signal-to-noise ratio is critical to distinguish IGM absorption from that arising from a DLA. Such an observing programme is worthwhile given the promise that GRBs have as probes of the epoch of reionisation." }, "0710/0710.4950.txt": { "abstract": "{} {\\par We study the forced rotation of Titan seen as a rigid body at the equilibrium Cassini state, involving the spin-orbit synchronization.} {\\par We used both the analytical and the numerical ways. We analytically determined the equilibrium positions and the frequencies of the 3 free librations around it, while a numerical integration associated to frequency analysis gave us a more synthetic, complete theory, where the free solution split from the forced one. } {\\par We find a mean obliquity of 2.2 arcmin and the fundamental frequencies of the free librations of about 2.0977, 167.4883, and 306.3360 years. Moreover, we bring out the main role played by Titan's inclination on its rotation, and we suspect a likely resonance involving Titan's wobble.} {} ", "introduction": "\\par Since the terrestrial observations of Lemmon et al. (\\cite{Lemmon93}), the rotation of Titan, Saturn's main satellite, has been assumed to be synchronous or nearly synchronous. This has been confirmed by Lemmon et al. (\\cite{Lemmon95}) and by Richardson et al. (\\cite{Richardson04}) with the help of Voyager I images. In this last work, Titan's rotation period is estimated at $15.9458 \\pm 0.0016$ days, whereas its orbital period is $15.945421 \\pm 0.000005$ days. \\par The spin-orbit synchronization of a natural satellite is very common in the solar system (such as for the Moon and the Galilean satellites of Jupiter) and is known as a Cassini state. This is an equilibrium state that has probably been reached after a deceleration of the spin of the involved body under dissipative effects, like tides. \\par Recently, Henrard and Schwanen (\\cite{Henrard04}) have given a 3-dimensional elaborated analytical model of the forced rotation of synchronous triaxial bodies, after studying the librations around the Cassini state. This model has been successfully applied by Henrard on the Galilean satellites Io (\\cite{Henrard05i}) and Europa (\\cite{Henrard05c}), seen as rigid bodies. Such studies require knowing some parameters of the gravitational field of the involved bodies, which cannot be considered as spheres. Another analytical study has been performed for Mercury by D'Hoedt and Lema\u00eetre (\\cite{DHoedt04}), for the case of a $3:2$ spin-orbit resonance. \\par Since the first fly-bys of Titan by the Cassini spacecraft, we have a first estimation of the useful parameters, more particularly Titan's $J_2$ and $C_{22}$ (Tortora et al. \\cite{Tortora06}), so a similar study of Titan's rotation can be made. In this paper, we propose a study of Titan's forced rotation, where Titan is seen as a rigid body. The originality of this study over Henrard's previous studies is that we use both the analytical and the numerical tools and compare our results. ", "conclusions": "\\par This paper offers a first study of Titan's rotation, where Titan is seen as a rigid body. We obtain a quasiperiodic decomposition of the forced solution, which can be split from the free solution in which Titan's obliquity plays an overwhelming role. Moreover, we find good matching between the frequencies of the free librations around the equilibrium, analytically and numerically evaluated, despite a model of circular orbit in the analytical study. However, we find a slight difference in the equilibrium obliquity. Finally, we cannot exclude a resonance between the proper mode $\\Phi_6$ and Titan's wobble. \\par The next fly-bys of Cassini spacecraft should give us more information on Titan's gravitational field, so we should be able to make a more accurate study on its rotation, that could include direct perturbations on the other Saturnian satellites. These perturbations are supposed to be small (see for instance Henrard \\cite{Henrard04}) and should be negligeable compared to the uncertainties we have on Titan's gravitational parameters. After that, the next step is to consider Titan as a multilayer non-rigid body and to study the consequences of its internal dissipation on the rotation." }, "0710/0710.3192_arXiv.txt": { "abstract": "M85 optical transient 2006-1 (M85\\,OT\\,2006-1) is the most luminous member of the small family of V838~Mon-like objects, whose nature is still a mystery. This event took place in the Virgo cluster of galaxies and peaked at an absolute magnitude of $M_{I}\\approx-13$. Here we present Hubble Space Telescope images of M85\\,OT\\,2006-1 and its environment, taken before and after the eruption, along with a spectrum of the host galaxy at the transient location. We find that the progenitor of M85\\,OT\\,2006-1 was not associated with any star forming region. The $g$ and $z$-band absolute magnitudes of the progenitor were fainter than about $-4$ and $-6$~mag, respectively. Therefore, we can set a lower limit of $\\sim50$\\,Myr on the age of the youngest stars at the location of the progenitor that corresponds to a mass of $<7$~M$_{\\odot}$. Previously published line indices suggest that M85 has a mean stellar age of $1.6\\pm0.3$~Gyr. If this mean age is representative of the progenitor of M85\\,OT\\,2006-1, then we can further constrain its mass to be less than $2$~M$_{\\odot}$. We compare the energetics and mass limit derived for the M85\\,OT\\,2006-1 progenitor with those expected from a simple model of violent stellar mergers. Combined with further modeling, these new clues may ultimately reveal the true nature of these puzzling events. ", "introduction": "\\label{Introduction} M85 Optical Transient 2006-1 (M85\\,OT\\,2006-1; J122523.82$+$181056.2) was discovered on 2006 Jan 6 by the Lick observatory supernova search team (Filippenko et al. 2001\\footnote{http://astro.berkeley.edu/$\\sim$bait/kait.html}) as a faint, $V\\sim19.3$~mag transient in the galaxy M85 (NGC\\,4382), which is at a distance of $17.8$~Mpc (Mei et al. 2007). Subsequent spectroscopy, as well as visible light and infra-red (IR) photometry, presented in Kulkarni et al. (2007), showed that M85\\,OT\\,2006-1 has a recession velocity of $880\\pm130$~km~s$^{-1}$, and is therefore associated with M85. Moreover, we showed that the temporal and spectral properties of this object are unlike those of supernovae, novae, or luminous blue variables. M85\\,OT\\,2006-1 peaked at absolute $I$-band magnitude of about $-13$. The light curve settled into a $\\sim60$~day plateau, followed by a decrease in bolometric luminosity during which the black-body emission peak shifted toward near-IR wavelengths. The early spectrum of M85\\,OT\\,2006-1, obtained six weeks after discovery, resembles that of a $\\sim4600$~K black body, with H$\\alpha$ and H$\\beta$ narrow emission lines (full width at half maximum of $\\sim350\\,$km\\,s$^{-1}$), along with several other unidentified emission lines. Spitzer IR observations obtained about six months after the discovery revealed a $\\sim1000\\,$K black body spectral energy distribution (Rau et al. 2007). The spectral and temporal properties of this object resemble those of M31-RV (discovered by Rich et al. 1989; e.g., Mould et al. 1990; Bryan \\& Royer 1992), V838~Mon (discovered by Brown 2002; e.g., Kimeswenger et al. 2002; Bond et al. 2003; Corradi \\& Munari 2007), and possibly the less studied object V4332~Sgr (Martini et al. 1999). However, the M85 transient is the most luminous member of the V838~Mon class. The favored model for this emerging class of V838~Mon-like objects (also known as luminous red novae\\footnote{this term was introduced by Kulkarni et al. 2007.}) is that they are the result of stellar mergers (e.g., Soker \\& Tylenda 2006). However, other models have been suggested to explain these objects (e.g., Retter \\& Marom 2003; Lawlor 2005). The nature of these events, with their energetics lying between the realms of supernovae and novae, remains uncertain. In this paper, we present Hubble Space Telescope (HST)- Advanced Camera for Surveys (ACS)/Wide Field Camera (WFC) and Near Infrared Camera and Multi-Object Spectrometer (NICMOS) observations, as well as Palomar 5\\,m spectroscopy, of the environment of M85\\,OT\\,2006-1. The observations are used to characterize the environment of the transient and to set a limit on the mass of the progenitor. ", "conclusions": "\\label{Disc} Although several models exist for V838~Mon-like objects (e.g., Soker \\& Tylenda 2003; Lawlor 2005), in the absence of detailed simulations, the nature of these objects remain elusive. A clue to their origin can be derived from their environment, luminosity function and rate. Given that only a small number of these objects are known, and they were found serendipitously in various searches, the luminosity function and rate are not well constrained. However, the fact that at least two events were observed in our Galaxy (i.e., V838~Mon and V4332~Sgr) in the last $\\sim13$ years suggests that they have a higher rate than SNe. We can set a lower limit on their rate, of $0.019\\,$yr$^{-1}\\,$L$_{MW}^{-1}$, at the $95\\%$ confidence level, where L$_{MW}$ is the Milky Way luminosity. Now we discuss the implications of our observations for a specific model for V838~Mon-like objects. Soker \\& Tylenda (2006) presented a model for violent stellar mergers in which, prior to the merger, the spins and orbital frequencies of the binary star are losing synchronization due to the Darwin instability (e.g., Eggleton \\& Kiseleva-Eggleton 2001). They found that for a given primary mass, the maximal energy production obtained for a binary mass ratio of $\\sim1/50$, is $\\sim2.5\\times10^{-3}GM_{1}^{2}/R_{1}$, where $G$ is the gravitational constant, and $M_{1}$ and $R_{1}$ are the mass and radius of the primary star. Given the upper limit on the progenitor mass, based on the mean stellar age in M85, $<2$~M$_{\\odot}$, and assuming a main-sequence mass-radius relation, $R\\propto M^{0.7}$, the maximum available energy in their model is short by a factor of three in the total energy production, as compared to the radiated energy of M85\\,OT\\,2006-1 in the first two months, $\\sim8\\times10^{46}$~ergs (assuming a distance of $17.8$~Mpc to M85; Mei et al. 2007). Moreover, it is expected that a large fraction of the energy will go into lifting the outer region of the star rather than radiated away. Furthermore, if the primary is an evolved star, then its radius will be larger than the radius of a main sequence star with the same mass, and the extracted energy will be even smaller. This suggests that either more detailed modeling of violent stellar mergers is required, or that this event is not the result of a violent stellar merger. Another possible solution is that the mass of the progenitor is somewhat larger. A larger progenitor mass will still be consistent with our upper limit of $7$~M$_{\\odot}$ which is based on the absence of stars brighter than $I\\sim-6$~mag. For example, according to Soker \\& Tylenda (2006) model, a $7$~M$_{\\odot}$ progenitor can yield $\\sim4$ times more energy than a $2$~M$_{\\odot}$ progenitor and may explain the discrepancy. We note, however, that other kinds of instabilities can lead to stellar mergers (e.g., in triple systems) and that the above comparison is valid only for the specific case discussed by Soker \\& Tylenda (2006). Existing hydrodynamical simulations of the common envelope phase in stellar mergers (and also star $+$ neutron star mergers) predict that the total dissipated energy is of the order of that observed in V838\\,Mon and M85\\,OT\\,2006-1 (e.g., Taam \\& Bodenheimer 1989; Terman et al. 1995; Terman \\& Taam 1996). Moreover, simulations of the common envelope phase predicts that most of the envelope will be ejected in the equatorial plane (e.g., Taam \\& Ricker 2006). Indeed, in Rau et al. (2007) we reported evidence suggesting that the expansion of M85\\,OT\\,2006-1 is asymmetric. However, more detailed hydrodynamical simulations of the vast parameter space available for stellar mergers are needed in order to understand these processes and to test if V838\\,Mon-like objects are indeed the results of stellar mergers. To summarize, we show that, in contrast to V838\\,Mon, but similarly to M31\\,RV, M85\\,OT\\,2006-1 was probably produced by members of an old stellar population ($>1\\,$Gyr), and that its progenitor/s mass was probably $\\ltorder2\\,$M$_{\\odot}$. These constraints narrow down the allowed venue of stellar models for the nature of this event." }, "0710/0710.4584_arXiv.txt": { "abstract": "We construct a physically motivated model for predicting the properties of the remnants of gaseous galaxy mergers, given the properties of the progenitors and the orbit. The model is calibrated using a large suite of SPH merger simulations. It implements generalized energy conservation while accounting for dissipative energy losses and star formation. The dissipative effects are evaluated from the initial gas fractions and from the orbital parameters via an ``impulse\" parameter, which characterizes the strength of the encounter. Given the progenitor properties, the model predicts the remnant stellar mass, half-mass radius, and velocity dispersion to an accuracy of 25\\%. The model is valid for both major and minor mergers. We provide an explicit recipe for semi-analytic models of galaxy formation. ", "introduction": "\\label{sec:intro} Major mergers between galaxies are central to the formation and evolution of elliptical galaxies \\citep{TT72,T77,MH94dsc}. The hierarchical buildup of galaxies in the $\\Lambda$CDM cosmology consists of a sequence of mergers, of which a significant fraction are ``major,\" involving progenitors with a mass ratio larger than 1:3. The gravitational interactions in such mergers have a dramatic effect on the dynamics and morphology of the galaxies, in particular turning rotating disks into pressure-supported spheroids. If the progenitors also contain gas, the mergers induce starbursts followed by gas consumption, which leads to aging stellar populations. The modeling of major mergers is therefore a key element in the attempts to confront the broad picture of galaxy formation with detailed observations. This is commonly performed via simulations incorporating Semi-Analytic Models (SAMs), where the complex physical processes are modeled using simplified parametric recipes. Advanced SAMs are currently attempting the non-trivial task of following the sizes and internal velocities of galaxies. For disk galaxies, sizes are evaluated using the halo virial radii $R_{\\rm vir}$ and spin parameters $\\lambda$ via $R_{\\rm disk} \\simeq \\lambda R_{\\rm vir}$, with some modifications due to the halo density profile (Fall \\& Efstathiou 1980; Mo, Mao \\& White; Bullock, Dekel et al. 2001; Dutton et al. 2006). The sizes of the remnants of gas-poor (``dry\") mergers, where the dominant interaction is gravitational, can be extracted from the properties of the progenitors and the orbital energy by assuming conservation of energy and relaxation to virial equilibrium \\citep{Cole00}. These considerations work well in simulations of relatively dry mergers, quite independently of the details of the orbit. While dry mergers may dominate in the formation of the most massive galaxies \\citep{Naab06}, the most common mergers are ``wet\" mergers of gaseous galaxies. It is thought that gas processes play an important role in the formation of ellipticals \\citep{RobertsonFP, Dekel06, Ciotti07}. As demonstrated below, the sizes predicted by dissipationless energy conservation can be off by a factor of a few for gas-rich mergers. Our goal is to construct a more accurate recipe to predict the size and velocity dispersion of the remnant of a wet merger given the properties of the progenitors and the orbital parameters. Cosmological mergers involve a complex mixture of variables (such as orbital parameters, gas fractions, mass ratio and bulge fraction) and physical processes (such as star formation and feedback), all of which can influence the properties of the merger remnants. Given a rich suite of high-resolution SPH merger simulations \\citep{thesis, Cox05}, that span the available parameter space and physical processes, albeit in a rather sparse and nonuniform manner, we seek a model that will properly represent the simulation results and enable an interpolation between them as well as an extrapolation to outside the simulated regime. For such a recipe to be successful, it should be based on a toy model that grasps the essence of the main physical processes involved in wet mergers. Our intuition is guided by the finding from the simulations that the remnants are more compact when the initial gas fraction is higher and when the first passage involves a stronger tidal impulse, namely when a larger fraction of the orbital energy turns into internal kinetic energy. At a first glance, this may seem surprising, as a system that gains energy is not expected to become more tightly bound, and indeed, the remnants of dry mergers are not very sensitive to the strength of the impulse. This dependence on gas fraction and on the impulse implies that a higher gas fraction and a stronger impulse are associated with a higher degree of dissipation, via shocks, collisions of gas clouds, and induced gas flows toward the centres of the merging systems. The resultant higher gas densities enhance the energy losses to radiation, leaving behind a more tightly bound remnant. In parallel, the higher degree of dissipation yields a stronger burst of star formation, which tends to be focused in the central region of the remnant. This understanding is the basis for our proposed recipe, which characterizes the merger by the impulse at first passage, evaluates the associated degree of dissipation and the resultant radiative energy losses and star formation, and accounts for these energy losses in the energy balance. A few free parameters with values of order unity can hopefully compensate for the crude approximations made. These approximations include, for example, an assumption of structural homology between the progenitors and remnant. The physically motivated recipe is then calibrated using the merger simulations, and its success is to be judged by its accuracy in matching the simulated remnant properties. In \\S \\ref{sec:methods} we describe the simulations used for this study. In \\S \\ref{sec:model} we present the details of our model for predicting remnant properties. In \\S \\ref{sec:Mass} we generalize the model to unequal mass mergers. In \\S \\ref{sec:Caveats} we discuss certain limitations of our model, and in \\S \\ref{sec:Conclusions} we summarize our conclusions. Appendix A discusses the details of our impulse approximation. Appendix B presents an explicit recipe for SAMs. ", "conclusions": "\\label{sec:Conclusions} We have developed a simple toy model for the physical processes involved in wet mergers of galaxies, and have calibrated it using a suite of hydrodynamical merger simulations. This modeling helped us to gain a better understanding of these processes, and provides a practical semi-analytic recipe for predicting post-merger galaxy properties in SAMs. Crude models of this sort have been used by \\citet{Cole00}, \\citet{Galics03}, and \\citet{Shen03}, but these models did not account for energy losses through dissipative processes, and they have not been calibrated against realistic merger simulations. Using a suite of merger simulations, we have demonstrated the key role of dissipative energy losses in determining the final radii and velocity dispersions of merger remnants. We found that the dissipative effects depend on the initial orbits of the progenitors. More violent, lower angular momentum orbits create greater disturbances in the gas disks, which in turn radiate more energy and produce more stars. This orbital ``violence'' can be parameterized through an impulse approximation for energy exchange between the orbital and internal components during the first close pass of the encounter. We present a physically-motivated, simulation-calibrated model that is capable of predicting star formation, central dark matter fraction, remnant radius and remnant velocity dispersion, given the properties of the progenitors and the initial orbital parameters of a merger. The non-dissipative energy conservation model often predicts radii that are off by a factor of $\\sim 2-3$, and it does not reproduce the spread due to orbital variations. Our model, which accounts for the dissipative energy losses, results in only $\\sim 25\\%$ errors in the predicted radius and velocity dispersion when a wide variety of progenitor types is considered. For a given progenitor type, the error in remnant properties is reduced to $\\sim 10\\%$, indicating that our model correctly captures the variation of remnant properties due to merger orbit. Since we used the whole available simulation suite to calibrate our model, via a few proportionality constants of order unity, a proper evaluation of the model performance is yet to be pursued using an independent suite of simulations." }, "0710/0710.1197_arXiv.txt": { "abstract": "% ", "introduction": " ", "conclusions": "" }, "0710/0710.4590_arXiv.txt": { "abstract": "The long, bright gamma-ray burst GRB~070125 was localized by the Interplanetary Network. We present light curves of the prompt gamma-ray emission as observed by Konus-WIND, RHESSI, Suzaku-WAM, and \\textit{Swift}-BAT. We detail the results of joint spectral fits with Konus and RHESSI data. The burst shows moderate hard-to-soft evolution in its multi-peaked emission over a period of about one minute. The total burst fluence as observed by Konus is $1.79 \\times 10^{-4}$~erg/cm$^2$ (20~keV--10~MeV). Using the spectroscopic redshift $z=1.548$, we find that the burst is consistent with the ``Amati'' $E_{peak,i}-E_{iso}$ correlation. Assuming a jet opening angle derived from broadband modeling of the burst afterglow, GRB~070125 is a significant outlier to the ``Ghirlanda'' $E_{peak,i}-E_\\gamma$ correlation. Its collimation-corrected energy release $E_\\gamma = 2.5 \\times 10^{52}$~ergs is the largest yet observed. ", "introduction": "The prompt gamma-ray emission of gamma-ray bursts (GRBs) is the most extensively studied aspect of these energetic explosions. Indeed, for twenty-five years after the discovery of GRBs \\citep{kleb73}, the prompt emission was the only GRB observable available. With the first afterglow observations at longer wavelengths \\citep{cost97,vanp97}, detailed analysis of burst models became possible. Presently, the \\textit{Swift} satellite is detecting $\\sim$ 100 bursts per year, most with rapid localization and followup. The exact mechanism which produces the prompt gamma-ray emission, with its characteristic smoothly broken power-law spectrum, has not been definitively established. Recent efforts to correlate burst observables with the intrinsic burst energetics have increased the importance of detailed spectral fitting for localized bursts \\citep[for a review, see][]{zhan07c}. Some correlations involve the peak spectral energy \\Ep, which is often above the $\\sim$150 keV cutoff of the \\textit{Swift} Burst Alert Telescope (BAT) passband. Several current observatories are capable of detailed spectral analysis of GRBs over the full range of \\Ep. Konus-W \\citep{apte95} is a double scintillator instrument on the WIND spacecraft. The Ramaty High Energy Solar Spectroscopic Imager (RHESSI) is a solar observatory which uses nine germanium detectors to image the Sun at X-ray to gamma-ray energies \\citep{lin02}. RHESSI's detectors are unshielded and receive emission from astrophysical sources like GRBs. The Wide-Band All-Sky Monitor (WAM) \\citep{yama05} aboard Suzaku is the large BGO anticoincidence shield for the Suzaku Hard X-Ray Detector. AGILE \\citep{trav06} and GLAST \\citep{ritz07} will give additional coverage at the energy range of \\Ep\\ and extend spectral coverage for GRBs up to tens of GeV. In this paper, we present Konus, RHESSI, and Suzaku observations of the bright GRB~070125. In Section \\ref{sec-obs}, we discuss the observations and the localization of the burst by the IPN. Section \\ref{sec-lc} contains the burst light curves, and in Section \\ref{sec-fits} we conduct joint spectral fits to the Konus and RHESSI data. ", "conclusions": "While GRB~070125 had a large measured prompt gamma-ray fluence, its spectral properties are unremarkable. The values of the best-fit spectral parameters are similar to those observed for other bright bursts \\citep[e.g.,][]{kane06}, and the spectral evolution observed is similarly common. The environment of GRB~070125 is unique, however \\citep{cenk08, chan08, updi08}, requiring a broad jet opening angle in broadband afterglow models \\citep{chan08}. After collimation correction, GRB~070125 has the most energetic prompt emission yet observed and is a significant outlier to the correlation between peak energy and $E_{\\gamma}$. GRB~070125 appears to weaken the claim that the Ghirlanda correlation has low dispersion. GRB~070125 is not a ``recognizable'' outlier to the Ghirlanda relation in the sense of \\citet{ghir07}, as it is highly consistent with the Amati relation. Its jet parameters have been derived from a rich and well-sampled afterglow dataset. While the circumburst environment of this GRB is unusually dense, this only highlights the assumption of a fairly narrow range of efficiency and density parameters for the majority of GRBs where broadband modeling of the afterglow has not been possible. The true dispersion of the correlation may in fact be larger. The physical significance of GRB spectrum--energy correlations has been questioned \\citep[e.g.][]{butl07,butl08}. In particular, detector trigger thresholds affect burst detection, and more complex selection effects govern the measurement of peak energies, redshifts, and afterglow breaks. These effects can influence the sample of GRBs with known redshift, $E_{peak,i}$, and $E_{\\gamma}$. \\citet{ghir08} examined the effect of trigger and spectral analysis thresholds in the \\Ep--fluence plane, finding that the \\textit{Swift}-detected burst sample was truncated by the spectral analysis threshold. Neither threshold truncated the pre-\\textit{Swift} burst sample. We were unable to confirm the source of the systematic shift in \\Ep\\ and fluence between the two instruments for this burst. Minor radiation damage was becoming noticeable in RHESSI detector 8 near the time of this work, mostly below the 65 keV cut utilized here. It is also possible that the Monte Carlo simulation of the RHESSI response is less accurate for such extreme off-axis angles, where a greater number of interactions with the cryostat may be expected. Our previous work had found excellent agreement in all fit parameters for independent RHESSI and Konus spectral fits for GRB 051103 and GRB 050717. For the short GRB 051103, Konus found $\\Ep = 1920 \\pm 400$~keV and a 20~keV--10~MeV fluence of $4.4 \\pm 0.5 \\times 10^{-5}$~ergs/cm$^2$ \\citep{kgcn051103, fred07}. A RHESSI fit yielded $\\Ep = 1930 \\pm 340$~keV and 20~keV--10~MeV fluence of $4.5 \\times 10^{-5}$~erg/cm$^2$ \\citep{bellhead06}. \\citet{krim06} found for a cutoff power-law fit to Konus data for GRB 050717 a best fit value of \\Ep $=2101 ^{+1934}_{-830}$~keV. A RHESSI fit to the same burst found \\Ep $=1550 ^{+510}_{-370}$~keV \\citep{wigg06}. Those bursts had RHESSI off-axis angles of 97 and 110 degrees, respectively. Joint spectral fits to \\textit{Swift}-BAT and RHESSI data for 25 bursts co-observed by the two instruments between December 2004 and December 2006 indicated that no offset in response normalization was needed for the two instruments \\citep{bellsf08}. However, for two of three bursts occurring during or after December 2006, the RHESSI data showed a significant deficit relative to \\textit{Swift}-BAT. The RHESSI polar angles for all three late bursts were between 90 and 110 degrees. These fits were conducted using only detectors 1 and 7, which do not appear to have radiation damage in background spectra during this interval. Nonetheless, these results suggest that the observed offset in the RHESSI and Konus fit parameters found here is more likely a consequence of increased radiation damage in the RHESSI detectors than a geometric effect or a generic offset in the RHESSI simulations. Future analysis of archival bursts may help identify the source of any systematic effects present here. It is clear, however, that joint fits between instruments capable of constraining the full range of \\Ep\\ are valuable in providing the most accurate and precise determination of the fit parameters." }, "0710/0710.4559_arXiv.txt": { "abstract": "We present BIMA observations of a 2$\\arcmin$ field in the northeastern spiral arm of M31. In this region we find six giant molecular clouds that have a mean diameter of 57$\\pm$13 pc, a mean velocity width of 6.5$\\pm$1.2 \\kms, and a mean molecular mass of 3.0 $\\pm$ 1.6 $\\times$ 10$^5$\\Msun. The peak brightness temperature of these clouds ranges from 1.6--4.2 K. We compare these clouds to clouds in M33 observed by \\citet{wilson90} using the OVRO millimeter array, and some cloud complexes in the Milky Way observed by \\cite{dame01} using the CfA 1.2m telescope. In order to properly compare the single dish data to the spatially filtered interferometric data, we project several well-known Milky Way complexes to the distance of Andromeda and simulate their observation with the BIMA interferometer. We compare the simulated Milky Way clouds with the M31 and M33 data using the same cloud identification and analysis technique and find no significant differences in the cloud properties in all three galaxies. Thus we conclude that previous claims of differences in the molecular cloud properties between these galaxies may have been due to differences in the choice of cloud identification techniques. With the upcoming CARMA array, individual molecular clouds may be studied in a variety of nearby galaxies. With ALMA, comprehensive GMC studies will be feasible at least as far as the Virgo cluster. With these data, comparative studies of molecular clouds across galactic disks of all types and between different galaxy disks will be possible. Our results emphasize that interferometric observations combined with the use of a consistent cloud identification and analysis technique will be essential for such forthcoming studies that will compare GMCs in the Local Group galaxies to galaxies in the Virgo cluster. ", "introduction": "In the inner disk of the Milky Way, the dominant component of the interstellar medium is molecular gas which is distributed in discrete cloud complexes \\citep{scoville87, scoville90}. Nearly all star formation in our Galaxy is associated with these clouds \\citep{scoville87, blitz93}. Therefore, understanding the distribution and properties of molecular clouds is a prerequisite for understanding galaxy evolution. Studies of the molecular gas emission in the Milky Way show that while the cloud mass spectrum extends over several orders of magnitude, a majority of the molecular gas mass is contained in large, massive cloud complexes, which are typically $\\sim$40 pc in diameter and have masses on the order of 1 $\\times$ 10$^5$ \\Msun; these complexes are usually referred to as giant molecular clouds (GMCs)\\footnote{In this paper, we will use the words giant molecular clouds, GMCs and cloud complexes interchangeably.} and are considered to be the basic organizational unit of the molecular interstellar medium \\citep{scoville90, combes91, young91}. Compared to the Milky Way, the properties of GMCs in external galaxies are not well known. The primary reason for this is a lack of high resolution, high sensitivity data which are necessary for spatially resolving individual complexes. Only in the Magellanic Clouds can single dish telescopes resolve extragalactic GMCs \\citep{israel93,rubio93, fukui01}. Even in the nearest spirals (e.g., M31 or M33), the angular size of a typical 40 pc GMC is $\\sim$12$\\arcsec$, approximately one half the resolution of the IRAM 30m single dish telescope at 2.6mm, the wavelength of the CO (J=1--0) line emission. For these galaxies, interferometric observations are necessary to resolve GMCs. GMCs have been detected using millimeter interferometers in the two nearest spiral galaxies, M33 \\citep{wilson90, engargiola03,rosolowsky07b} and M31 \\citep{vogel87,wilson93,loinard98,rosolowsky07a}. In M33, \\citet{wilson90} observed 19 different fields and identified over 30 GMCs but only 9 GMCs were suitable for a study of cloud properties. The number of GMCs detected in most spiral galaxies has been less than a dozen, presumably due to the small coverage area. This is beginning to change. Recently a survey of M33 was undertaken by the BIMA array. The angular resolution was coarse ($\\sim$13$\\arcsec$) but adequate to identify nearly 150 GMCs (\\citealt{engargiola03}, see also \\citealt{rosolowsky07b}). With the upcoming CARMA array, ground-breaking studies like the M33 survey are likely to become routine for external galaxies. It is important to note that previous studies of GMCs in external galaxies have used different instruments with varying angular and linear resolutions and varying sensitivity. Moreover, these studies have used different cloud-identification methods to define GMCs and determine their properties. Yet, in almost all of these studies, the observations have detected discrete molecular features similar to Galactic GMCs. Particularly notable is the consistent linewidth-size relationship in most of these observations. However, several studies have noted differences between Galactic GMCs and those found in some irregular and dwarf galaxies. In the irregular galaxies NGC 6822 \\citep{wilson94a} and the SMC \\citep{rubio93}, the clouds are slightly less massive, and perhaps smaller than clouds seen in the Milky Way. In the LMC, \\citet{cohen88} find that for a given size and line width, clouds are about six times fainter in CO than comparable clouds in the Milky Way. The same is true of the dwarf NGC 1569 where the CO-to-H$_2$ conversion factor is reported to be six times larger than the Galactic value \\citep{taylor99}. In IC10, the cloud sizes and masses are similar to the Milky Way clouds, but the CO-to-H$_2$ conversion factor may be twice as high as that seen in the Galaxy if the clouds are assumed to be self-gravitating \\citep{wilson91}. These studies suggest that there are noticeable differences between the Galactic GMCs and those found in irregular or dwarf galaxies, that may be attributed to the different host environments (e.g., metallicity) of these galaxies. GMCs in the three nearest spiral galaxies, M31, M33 and the Milky Way can be observed at sufficiently high resolution ($\\sim$20 pc) to remove ambiguities from coarse resolution. Are the GMC properties in these spiral galaxies fundamentally different? \\citet{wilson90} claim that no GMC larger than 4$\\times$10$^5$ \\Msun is seen in M33, and unlike the Milky Way, 50\\% of the molecular mass in M33 resides in small clouds with masses less than 8$\\times$10$^4$ \\Msun, based on a comparison of interferometric and single dish flux from the CO(J=1--0) line. On the other hand, analysis of the $^{13}$CO/$^{12}$CO line ratio for a number of M33 clouds suggests that the total mass measured from the $^{12}$CO (J=1-0) line may be overestimated and the actual conversion factor from the CO to H$_2$ mass may be lower in the diffuse gas; in that case, the amount of mass in GMCs would constitute most of the molecular mass in M33 \\citep{wilson94b}. With coarser resolution data, \\citet{engargiola03} find a steep GMC mass spectrum in M33 and report a characteristic mass of 7 $\\times$ 10$^4$\\Msun. They conclude that molecular clouds in M33 are formed rapidly from the atomic gas and are short lived entities. On the other hand, in M31 \\citet{vogel87} and \\citet{wilson93} note that the four GMCs they detect are similar to Galactic GMCs. \\citet{loinard99} smooth a section of the Carina arm in the Milky Way to the resolution of single dish M31 maps and conclude that there are general similarities between the distribution of molecular gas in M31 and the Milky Way. Similar results were also found by \\citet{heyer00} who compared the molecular gas in M31 and the Milky Way using single dish observations; in their study,\\citet{heyer00} find a number of similarities in the molecular gas proeprties between M31 and the Milky Way such as the amount of molecular gas at radii larger than 8 kpc, large arm to interarm contrasts, and similar surface brightness, line widths and spacings of GMCs in spiral arms. \\citet{rosolowsky07a} has recently completed a large survey of clouds in M31. He also finds that the GMCs in M31 are similar to those found in the Milky Way. These results depend on comparison of properties derived from datasets which are obtained with a different instrument and analyzed with different methods of cloud identification. The Milky Way data, for example, come from single dish observations with a signal to noise that is at least 2-4$\\times$ higher than in the M33 and M31 datasets. The datasets also differ in that single-dish telescopes are inherently different than the aperture synthesis maps: the interferometer will miss smooth, extended emission whereas the single-dish data may suffer from beam dilution. The Milky Way survey also suffers from varying linear resolution. Finally, the methods for identifying clouds in the different Milky Way studies vary. These studies (e.g., \\citealt{dame86,sanders85}) have usually projected a typical (l,b,v) cube on one of its axes and used a integrated intensity contour to identify discrete features they call clouds or complexes. The sizes were measured from the total area of the cloud, assuming that the cloud was spherical \\citep{dame86}, or by measuring the chord across the velocity centroid in a position-velocity diagram \\citep{sanders85}. In M33, \\citep{wilson90} used a two-tiered selection criterion which required that a cloud be 3$\\sigma$ or more in 2 adjacent channels and that its summed flux be at least 3$\\sigma$ above the noise. The lack of consistency in these methods and differences in observing techniques are worrisome. A proper comparative study of GMCs in these galaxies requires that the differences described above be eliminated. This is particularly important in view of future telescopes such as CARMA and ALMA which will have the ability to mosaic large regions of galaxies like M31 and M33. The large scale surveys will have the detail necessary for in-depth comparisons of the molecular ISM in Local Group and more distant galaxies. These future studies will address the fundamental question of how the molecular ISM varies from galaxy to galaxy. Knowing this is critical to understanding star formation and galaxy evolution. In this paper, we make an attempt to do a fair comparison of the molecular emission in the three nearest spiral galaxies, i.e., M31, Milky Way and M33. We compare the M31 complexes to the simulated Milky Way complexes and previous surveys, using a cloud-identification technique in which we identify complexes using a single integrated intensity contour (\\S \\ref{newres}), and discuss how previous conclusions about M33 would change if the same technique of cloud identification were used (\\S \\ref{just}). An important addition to this analysis would be the addition of single dish data for the M31 and M33 clouds which would allow for an even more comprehensive comparison. Of course, one would still need to project the Milky Way clouds to the distance of Andromeda to compare the clouds at the same spatial resolution. Our method is intended to lay the groundwork for future studies that will compare molecular clouds across galactic environments and across different galaxy types with telescopes such as CARMA, SMA and ALMA. In \\S \\ref{methods}, we describe two automated cloud identification algorithms (Gaussclump developed by \\citet{stutzki90}, and Clumpfind developed by \\citet{williams94}), and their advantages and disadvantages. While we do not advocate using these algorithms to describe the characteristics of GMC populations in a galaxy, these algorithms may be used to compare the M31, M33 and the simulated Milky Way data. We describe the results of this experiment in \\S \\ref{auto} and discuss our overall conclusions in \\S \\ref{conc}. Preliminary results from this work have been published in \\citet{sheth00}. ", "conclusions": "\\label{conc} In the northeastern spiral arm of M31, we have identified six distinct, large complexes of molecular gas. All of these complexes lie along the spiral arm dust lanes, or in one case (Complex E), along a dust spur in between the dust lanes. These complexes have a mean diameter of 57$\\pm$13 pc, a mean velocity width of 6.5$\\pm$1.2 \\kms, and a mean molecular mass of 3.0$\\pm$1.6 $\\times$ 10$^5$ \\Msun and are indistinguishable from those found in the Milky Way e.g., \\citet{dame86}. Meaningful comparison of GMCs in different galaxies requires consistent analysis of data taken with different instruments and different cloud identification techniques. This paper represents an attempt at eliminating such differences and comparing GMCs in M31, M33 and the Milky Way in as consistent a manner as possible. We have simulated three Milky Way complexes at the distance of M31 and observed them with the BIMA array. The simulations show that interferometers are excellent at recovering molecular complexes in the Local Group galaxies. We compared the simulated Milky Way complexes to the M31 data using an integrated intensity contour method and found that the complexes thus identified and analyzed fell on the same linewidth size relationship as found by \\citep{dame86}. The M33 clouds if analyzed in this manner also yield larger clouds than previously stated. Larger interferometric surveys of these galaxies are necessary to compare the GMC distribution and mass function to that in the Milky Way (e.g., \\citealt{engargiola03, rosolowsky07a}). Finally, we compared two automated algorithms, Clumpfind and Gaussclump, and found that Gaussclump was better at recovering cloud properties than Clumpfind for low dynamic range data. Our experiments revealed that such algorithms are resolution dependent and the inherent clumpy nature of molecular emission prevents these methods from identifying clouds larger than one or two resolution elements. Hence we caution against using these methods to characterize GMC populations in galaxies. Still, these may be used to quickly compare properties of clouds in data with similar noise and resolution characteristics. Using Gaussclump, we find that the cloud properties in M31, M33 and simulated Milky Way data are indistinguishable." }, "0710/0710.4135_arXiv.txt": { "abstract": "We present an analysis of the optical colors of 413 Virgo cluster early-type dwarf galaxies (dEs), based on Sloan Digital Sky Survey imaging data. Our study comprises (1) a comparison of the color-magnitude relation (CMR) of the different dE subclasses that we identified in Paper III of this series, (2) a comparison of the shape of the CMR in low and high-density regions, (3) an analysis of the scatter of the CMR, and (4) an interpretation of the observed colors with ages and metallicities from population synthesis models. We find that the CMRs of nucleated (\\,dE(N)\\,) and non-nucleated dEs (\\,dE(nN)\\,) are significantly different from each other, with similar colors at fainter magnitudes ($\\mr\\gtrsim 17$ mag), but increasingly redder colors of the dE(N)s at brighter magnitudes. We interpret this with older ages and/or higher metallicities of the brighter dE(N)s. The dEs with disk features have similar colors as the dE(N)s and seem to be only slightly younger and/or less metal-rich on average. Furthermore, we find a small but significant dependence of the CMR on local projected galaxy number density, consistently seen in all of $u-r$, $g-r$, and $g-i$, and weakly $i-z$. We deduce that a significant intrinsic color scatter of the CMR is present, even when allowing for a distance spread of our galaxies. No increase of the CMR scatter at fainter magnitudes is observed down to $\\mr \\approx 17$ mag ($\\Mr \\approx -14$ mag). The color residuals, i.e., the offsets of the data points from the linear fit to the CMR, are clearly correlated with each other in all colors for the dE(N)s and for the full dE sample, implying that, at a given magnitude, a galaxy with an older stellar population than average typically also exhibits a higher metallicity than average. Given the observational data for Virgo dEs presented here and in the previous papers of this series, we conclude that there must be at least two different formation channels for early-type dwarfs in order to explain the heterogeneity of this class of galaxy. ", "introduction": "\\label{sec:int} It has long been known that a close correlation exists between the colors and luminosities of early-type galaxies, in the sense that more luminous objects have redder colors \\citep[e.g.,][]{bau59,deV61,fab73,sanvis78a,cal83}. A striking observation is the universality of the color-magnitude relation (CMR): it was found to be equal, within the measurement errors, for E and S0 galaxies within clusters, groups, or the field \\citep{fab73,sanvis78a,sanvis78b,bow92}, leading \\citet{fab73} to state that the colors of elliptical galaxies ``are independent of all physical properties studied other than luminosity''. \\citeauthor{fab73} also showed that a similar relation exists between the strength of spectral absorption features and luminosity, which basically is the spectroscopic analogue to the CMR. She interpreted the CMR as a trend of increasing metallicity with luminosity, which today is still considered to be the primary determinant of the CMR \\citep[e.g.,][]{kod97,cha06}. This can be understood with a higher binding energy per unit mass of gas in more massive galaxies, leading to stronger enrichment of the stellar populations. Recently, \\citet{ber03IV} showed that color seems to correlate even more strongly with velocity dispersion than it does with luminosity. This implies that the CMR itself is most likely just a combination of the relation of luminosity and velocity dispersion \\citep[``Faber-Jackson relation'',][]{fab76} and that of color and velocity dispersion. As \\citet{mat05} point out, the latter might hint at a more fundamental relation, namely between galaxy metallicity and mass. While this trend includes the dEs, there has been disagreement about whether or not they follow the same CMR as the giant ellipticals. \\citet{deV61} found the dwarfs to be ``systematically bluer'' than the giants (but nevertheless following a CMR), whereas \\citet{cal83} reported a linear CMR over a range of $-15 \\le \\MV \\le -23$ mag. However, his Figure~3 actually suggests that the slope of the CMR might indeed be slightly different for the dEs, in the same way as the results of \\citet{deV61} suggested (i.e., with decreasing magnitude, dwarfs become bluer more rapidly than giants do). \\cite{mat05} found a change in the slope of the Faber-Jackson relation for ``faint early-type'' galaxies --- here, ``faint'' means $-17.3 \\le \\Mb \\le -20.5$ mag, thus reaching only slightly into the dwarf regime. \\citet{der05} found an even larger difference in slope for a sample of 15 dEs, but argued that this is consistent with theoretical models, due to the dynamical response to starburst-induced mass loss, which is stronger for objects of lower mass. Significant constraints to such models could only be provided with a better understanding of dE formation. In contrast to giant ellipticals, formation mechanisms proposed for dEs in clusters are typically not based on an early formation epoch, but rather, on infall and subsequent transformation of late-type galaxies through gas-stripping and tidally enhanced star formation \\citep[e.g.,][]{dav88,moo96,vZe04a,sab05,del07}. After having established a subdivision scheme of Virgo cluster early-type dwarf (dE) galaxies into subclasses with different shapes and distributions \\citep[Paper III of this series]{p3}, we can now proceed to the next logical step, namely to exploiting the wealth of data provided by the Sloan Digital Sky Survey (SDSS) Data Release 5 \\citep[DR5,][]{sdssdr5} by a multicolor analysis of our sample of 413 dEs. However, the colors of different dEs, or of dEs of different subclasses, can not straightforwardly be compared with each other: the existence of the CMR requires such a comparison to be done either at fixed magnitude or with a correction for magnitude differences. For this reason, we shall explore the dE colors mainly through an analysis of their CMR. One would naively expect that, if two given dE subclasses formed through different mechanisms, their resulting CMRs should display differences as well, since the relation of galaxy mass to the properties of its stellar population should depend to some extent on how and when the latter was formed. On the other hand, the apparent universality of the CMR -- mainly defined for giant ellipticals -- would seem to argue against such differences. \\citet{conIII} found considerable scatter of the CMR of dEs in the Perseus cluster at magnitudes $\\Mb \\ge -15$ mag, apparently caused by two different sequences of dwarfs in color-magnitude space. They argue that the early-type dwarfs must have multiple origins --- similar to the conclusion of \\citet{pog01} about an apparent bimodal metallicity distribution of Coma cluster dwarfs that might hint at more than one formation channel. We should be able to test the presence of such a bimodality in more detail, given our ``preparatory work'', namely the separation of dE subclasses that have different shapes and distributions (Paper III). Moreover, the SDSS multicolor data enable us to construct CMRs in more than one color, and also to analyze our galaxies in color-color space, thereby translating colors into ages and metallicities. Since \\citet{rak04} found the dE(nN)s in the Coma and Fornax clusters to be younger and to have a higher metallicity than the dE(N)s, we can perform a similar analysis for our Virgo cluster galaxies, allowing us to test the similarity of dE populations of different clusters. ", "conclusions": "\\label{sec:discussion} We have analyzed the colors of 413 Virgo cluster dEs by constructing color-magnitude relations (CMRs) for different dE subclasses and different local densities, as well as by comparing them to theoretical colors from population synthesis models of \\citet{bc03}. We found significant differences between the CMRs of dE(N)s and dE(nN)s, as well as between the CMRs at low and high local projected densities. The models imply that the brighter dE(nN)s are younger than the dE(N)s and/or have lower metallicities. The dE(di)s are more similar to the dE(N)s, yet still seem to be slightly younger and/or less metal-rich on average. A significant intrinsic color scatter of the CMR is present. The color residuals about the CMR are correlated between different colors for the dE(N)s, and partly also for the dE(nN)s, such that a galaxy falling on the blue side of the CMR in one color also does so in the other color. We find no increase in the color scatter at fainter magnitudes down to $\\mr \\approx 17$ mag ($\\Mr \\approx -14$ mag). The dE(N)s are consistent with having a nearly ``perfect'' intrinsic correlation of colors, i.e., if the intrinsic $u-r$ color of a dE(N) lies on the red side of the respective CMR, the same is true in almost all cases for the intrinsic $i-z$ color. This is particularly interesting, since $u-r$ is more sensitive to the age of the stellar population, while $i-z$ is sensitive to metallicity. A simple, straightforward interpretation is that when the stars in a dE are on average older than the typical value at that dE's luminosity, then they are also more metal rich, and vice versa. Assuming a direct correlation between luminosity and galaxy mass, we can speculate that the intrinsic scatter of the CMR could, for a given initial mass, reflect a spread in star formation rate or in the efficiency with which gas is turned into stars, perhaps caused by environmental effects. Neglecting other possible effects, a higher SFR at a given initial (gas) mass would lead to stronger enrichment, i.e.\\ to a higher metallicity, than a lower SFR. The gas would be consumed more rapidly, thus reaching the end of star formation earlier than with a lower SFR, and consequently yielding older stars on average. \\subsection{Subclass colors and stellar populations} \\label{sec:sub_discuss1} While we found in Section~\\ref{sec:subclasses} that dE(nN)s and dE(N)s follow different CMRs, we then discovered in Section~\\ref{sec:density} that the CMR depends on environmental density. Since dE(nN)s and dE(N)s populate different density regimes (Paper III), we should compare the CMR of the dE(nN)s to that of the \\emph{low-density} subsample of dE(N)s: the median density of the latter is 1.18 (in units of the logarithm of the number of galaxies per square degree), and it is 1.20 for the dE(nN)s. In contrast, the median density of the full sample of dE(N)s is 1.37. However, the CMR of the low-density dE(N)s is still significantly different from that of the dE(nN)s in $g-r$ and $g-i$; the probability for a common distribution is $\\le$0.1\\% for the half-light aperture. Likewise, the color difference between dE(N)s and dE(nN)s at the bright reference magnitude is $0.09$ mag in $u-r$ and $0.05$ mag in both $g-r$ and $g-i$, using the half-light aperture. This difference only changes by $0.01$ mag in $u-r$ and $g-i$ and even less in $g-r$ when considering only the low-density dE(N)s. Thus, the colors of the (bright) dE(nN)s do differ from those of the dE(N)s even if we allow for the different sampling in density. The same test can be done for the dE(di)s: their densities (median value 1.18) are also much more comparable to those of the low-density dE(N)s than to those of the full sample. While the color difference between the dE(di)s and the full dE(N) sample at the bright reference magnitude is only $0.02$ mag in $u-r$, $0.03$ mag in $g-r$, and $0.04$ mag in $g-i$, the statistical comparison of their CMRs yields significant differences (Section~\\ref{sec:subclasses}). This changes when we compare the dE(di)s to only the low-density dE(N) sample. Although the colors of the dE(di)s are still slightly bluer than those of the low-density dE(N)s, none of these differences is statistically significant. Whether or not this could indicate a close relation between dE(di)s and dE(N)s despite their very different shapes will be discussed in Section~\\ref{sec:sub_discuss2}. \\citet{rak04} determined ages and metallicities for 91 dEs in the Coma and Fornax clusters, based on narrowband photometry. They derived ages above 8 Gyr for the dE(N)s, which they found to be about 5 Gyr older than the dE(nN)s. At least qualitatively, and in a relative sense, this would be consistent with our results. As for the metallicities, \\citeauthor{rak04} found the dE(N)s to have \\emph{lower} metallicities than the dE(nN)s, conjecturing that ``globular clusters and dEN galaxies are primordial and have metallicities set by external constraints such as the enrichment of their formation clouds.'' This is not in agreement with our results for the brighter magnitudes: there, we find the metallicities of the dE(N)s to be either similar (in the inner part of the galaxies, see Figure~\\ref{fig:modelsplusdata}) or higher (Figure~\\ref{fig:modelsplusdata_metal}) than those of the dE(nN)s. Due to the fact that the Virgo cluster is a dynamically less relaxed structure than the Coma and Fornax clusters, it would be particularly interesting to find systematic differences in the stellar content of their galaxy populations. However, it would be premature to conclude that the dE(nN)s and dE(N)s in Virgo behave inversely to those in the other clusters --- the sample of \\citet{rak04} only comprises 10 dE(nN)s of the Coma cluster and 9 dE(nN)s of the Fornax cluster, and moreover, the dE(nN)s show a considerable color scatter in the metallicity-sensitive color $i-z$. Whether or not these issues can account for the differences needs to be analyzed in future work. The question of whether our fitted CMRs would be consistent with being mainly a luminosity-metallicity relation can be qualitatively addressed by comparing the respective CMR values at the bright and faint reference magnitude in Figure~\\ref{fig:modelsplusdata_metal}. Overall consistency with this interpretation is present for all subclasses, but we cannot exclude the additional presence of at least some age differences with magnitude. A general problem for the interpretation of the colors also is that the color values for some data points lie at ``too large'' an age, i.e., they fall above the age of the Universe for our model tracks. We thus need to emphasize again that many simplifications entered the calculation of these tracks, like the fact that the models are calculated at a fixed metallicity, or the rather simple star formation histories that we consider. All our interpretations of observed dE colors are always done within this simplified framework of stellar population models. \\subsection{The minimum number of formation scenarios} \\label{sec:sub_discuss2} On the basis of our observational analyses presented here and in the previous papers of this series, we now attempt to answer the question of how many different dE formation mechanisms there must be \\emph{at least} in order to explain the diversity of dE subclasses. To start with, how confident can we be that the (flat) dE(di)s do not belong to the same intrinsic (sub-)class as the (round) dE(N)s? The distribution of intrinsic shapes of the brighter dE(N)s, as deduced in Paper III, is rather broad and includes a significant number of flat objects, even down to axial ratios of 0.3. Most of the range of intrinsic axial ratios of the dE(di)s is thus covered by the range of values of the dE(N)s. Moreover, 73\\% of the dE(di)s are nucleated. While their colors are somewhat different from the full sample of dE(N)s, the difference is not anymore significant when compared only to the low-density dE(N)s, which have the same median density as the dE(di)s (Section~\\ref{sec:sub_discuss1}). If disk substructure, like spiral arms or bars, could only occur in the flattest dEs, due to, e.g., the kinematical configuration of these objects, we would have automatically selected only intrinsically flat galaxies in our search for disk features (Paper I), and would obviously have found their flattening distribution to be consistent with disk galaxies. The fact that these show no central clustering could then be explained, for example, by the much stronger tidal heating that a galaxy experiences in denser regions of the cluster, leading to an earlier destruction of disk features \\citep[cf.][]{mas05}. Similarly, if dE(N)s and dE(di)s originated from a morphological transformation of infalling late-type spirals through galaxy harassment \\citep{moo96}, one could imagine that the amount of transformation depended on how close the encounters with massive galaxies were that led to it --- and the probability for close(r) encounters is obviously higher in the cluster center, leading to rounder objects without disk features. Why, then, are there almost no fainter dE(di)s? While we concluded in Paper I that we most likely missed a significant amount of dE(di)s at fainter magnitudes due to our detection limits, we also argued that the true number fraction \\emph{does} decrease when going to fainter objects. As a possible explanation, we can speculate that disk substructure might be more likely to occur in more massive galaxies, possibly connected to the presence of a certain amount of rotational velocity. These qualitative considerations demonstrate that dE(di)s and dE(N)s could, in principle, have formed through the same formation process, keeping our counter of necessary dE formation mechanisms at 1 for the moment. Could the dE(nN)s be also related to the dE(N)s and be formed by the same process? At least for the bright subsamples, the colors of dE(nN)s and dE(N)s differ significantly, indicating younger average stellar ages of the dE(nN)s, or lower metallicities, or both (Section~\\ref{sec:stelpop}). This still holds true even when the different density distributions of dE(nN)s and dE(N)s are accounted for (Section~\\ref{sec:sub_discuss1}). One might thus conjecture that the dE(nN)s simply formed more recently. The significantly flatter shapes of the bright dE(nN)s could then possibly be explained in the sense that they still need to experience several (further) encounters with massive galaxies, leading to further morphological transformation. One would, though, need to invoke another assumption, namely that the faint dE(nN)s, which are already significantly rounder than the bright dE(nN)s, were much more affected by the first tidal encounters, while the bright dE(nN)s were able to partly preserve their initial shape\\footnote[7]{~However, note that the bright dE(nN)s might not even have experienced any tidal encounters yet, but could have been born as thick, puffy systems \\citep{kau07}.} --- this could possibly be explained by the difference in mass. Another requirement of this scenario would be that the distribution of dE(nN)s within the cluster, which is not centrally concentrated at all (Paper III), would need to shift towards significantly larger local densities within the next few Gigayears, in order to be similar to the distribution of today's dE(N)s. However, \\citet{conI} derived a two-body relaxation time for the Virgo dEs of much more than a Hubble time. Even violent relaxation, which probably only applies for the case of infalling or merging groups, would take at least a few cluster crossing times $t_{\\rm cr}$, with $t_{\\rm cr}\\approx 1.7$ Gyr for Virgo \\citep{bos06}. It thus seems difficult to reconcile these numbers with the required dynamical process. A further, obvious point is that nuclei would need to form soon in the dE(nN)s. Perhaps nuclei are currently being formed in the centers of the dE(bc)s where we are witnessing ongoing star formation (Paper II). Such a scenario would lead to nuclei whose stars were clearly younger than the vast majority of their host galaxies' stars. This would, however, be hardly consistent with the results of \\citet{acsvcs8}, who found \"old to intermediate-age populations\" of dE nuclei, and of \\citet{lotz04}, who measured similar colors of nuclei and dE globular clusters. An alternative would be that most nuclei form through coalescence of globular clusters \\citep[e.g.][]{oh00}. Yet in neither case do we find an explanation for why the ratio of nucleated and non-nucleated dEs should increase strongly with luminosity \\citep{san85b} if the latter were the immediate progenitors of the former. Taken together, our observational results do not allow to explain dE formation with less than two different processes. We thus conclude by repeating the statement of \\citet{vZe04a}, ``we caution against single-channel evolutionary scenarios.'' Early-type dwarfs are not a homogeneous class of objects, and we strongly recommend to separately analyze the properties of dEs belonging to different subclasses in any future study of dEs." }, "0710/0710.0699_arXiv.txt": { "abstract": "The recent discovery of super-Earths (masses $\\leq$ 10 $M_{\\oplus}$) has initiated a discussion about conditions for habitable worlds. Among these is the mode of convection, which influences a planet's thermal evolution and surface conditions. On Earth, plate tectonics has been proposed as a necessary condition for life. Here we show, that super-Earths will also have plate tectonics. We demonstrate that as planetary mass increases, the shear stress available to overcome resistance to plate motion increases while the plate thickness decreases, thereby enhancing plate weakness. These effects contribute favorably to the subduction of the lithosphere, an essential component of plate tectonics. Moreover, uncertainties in achieving plate tectonics in the one earth-mass regime disappear as mass increases: super-Earths, even if dry, will exhibit plate tectonic behaviour. ", "introduction": "Until recently, Earth was the largest terrestrial object known to exist. However, five super-Earth planets (a class defined as having a mass between 1-10 $M_{\\oplus}$- earth-masses) have been detected in the last few years \\citep{Rivera_et_al:2005,OGLE-5.5:2006,Lovis_et_al:2006,Udry_et_al:2007}. The five planets have masses in the 5-10 $M_{\\oplus}$ range, but we do not have information on their sizes and cannot be sure if these are really rocky terrestrial planets. However, their discovery provides some evidence that super-Earths might be common and it is only a matter of chance that our Solar System has none. Some of these planets might be in the 'habitable zone', where the radiation from the star allows for the presence of liquid water, but only their thermal and chemical evolution will determine if they are, in fact, habitable. In turn, their thermal evolution and surface conditions depend on and affect their tectonic regime. Currently, Earth is the only planet where plate tectonics is active. Furthermore, this mode of convection has dominated our planet's geological history, is associated with geochemical cycles and thus, has been proposed as a required mechanism for life on Earth \\citep{Walker_et_al:1981}. Here we address whether or not super-Earths are likely to have plate tectonics or be in a stagnant lid convection like Mercury and Mars. ", "conclusions": "In summary, convection is more vigourous in massive terrestrial planets, making their lithospheres thinner and therefore reducing lithospheric strength. Furthermore, they achieve larger stresses owing primarily to larger velocities and therefore can more easily overcome the lithospheric resistance to deformation. Plates may reach negative buoyancy on super-Earths despite their relative younger ages. This scenario is suitable for the failure of the plate and subsequent subduction, which is a necessary step for plate tectonics. Given that Earth's convective state leads to plate tectonics, the more favorable conditions experienced by super-Earths will inevitably lead to plate tectonics. Furthermore, planets of similar mass should have the same potential to exhibit plate tectonics. Conversely, this physics can help explain why small planets like Mars, Mercury and the Moon do not exhibit plate tectonics. \\subsection{Role of Water} Venus is only slightly smaller than Earth and does not exhibit plate tectonics, although some authors \\citep{Turcotte:1993,Jellenik_water:2005J} have suggested it may have in the past. This observation indicates that the $\\sim$1 earth-mass case falls within a zone of transition between `hard' stagnant lid and mobile plate regimes. In this case, characteristics other than M may be important to the dynamics of the lithosphere. For example, the high surface temperature of Venus might lead to a weak, highly deformable boundary layer that would not support the coherent plate-like behaviour that characterizes oceanic plates on Earth. Moreover, plate strength is relatively large compared to the mantle driving force in the one earth-mass case; yield stresses are on the order of 1-5 GPa for olivine (the representative upper mantle mineral) \\citep{Chen_et_al:1998}, whereas our calculations, in agreement with more detailed models \\citep{Becker_Oconnell:2001}, suggest an underlying driving force of only 10 MPa. Since slip can occur on pre-existing faults at stress values of a few MPa, the existence of plate tectonics in the one earth-mass regime may thus depend crucially on the conditions required to initiate subduction. The presence of water is one possible mechanism to reduce the yield strength of a plate and friction on faults. Experiments show that water reduces the yield strength of olivine by 62\\% when raising the temperature from 25 to 400$^{\\circ}$C at 10 GPa, compared to a drop of 39\\% in dry olivine \\citep{Chen_et_al:1998}. Hence, the hydration level of Venus' mantle, which is 1-2 orders of magnitude lower than on Earth \\citep{Zolotov_et_al:1997}, may make it very difficult for convective forces within this planet to overcome plate resistance. For larger planets, super-Earths, these issues become less relevant. A wet super-Earth will clearly have enough driving force to sustain subduction. But, more importantly, the consequences for initiating subduction associated with the hydration of a one earth-mass planet (i.e., a reduction of the yield strength by half) would be similar to a doubling of the mass of the planet (Fig. 2). That is to say, both scenarios would be as likely to initiate and maintain subduction. \\subsection{Atmospheric Observables} The difference between a Super-Earth with active plate tectonics and one with stagnant lid is in the access of upper mantle material and gasses to the atmosphere. The first case allows several global geochemical cycles to operate, like the CO$_2$ and SO$_2$ ones. For example, cases in our Solar System comprise: Earth with a CO$_2$ cycle and possibly early Mars with a SO$_2$ cycle \\citep{Halevy_et_al:2007}. Earth has had stable modest levels of atmospheric CO$_2$ (between 160-7000 ppm -- \\citet{Royer_et_al:2001}) in the last 0.5 Gy whereas Venus\u2019 levels stand today at 96\\%. A planet with plate tectonism and a carbonate rock reservoir has an efficient built-in cycle that stabilizes climate at temperatures within the liquid water regime \\citep{Kasting:1996}. A super-Earth that has plate tectonics and weathering capabilities can be expected to have CO$_2$ atmospheric concentrations that would yield temperatures around liquid water. Therefore, evidence against the presence of plate tectonics on an exoplanet would be the detection of high values of CO$_2$ for the age of the star, type of star and orbital distance. An SO$_2$ based atmosphere is also possible and the same reasoning would apply, since the sulfur cycle operates analogously to the carbon cycle. But obviously, more theoretical research is necessary to model the details and predict the right observable signatures. In conclusion, we show here that as mass increases, the process of subduction, and hence plate tectonics, becomes easier. Therefore, massive super-Earths will very likely exhibit plate tectonics. In the future with TPF by NASA and Darwin by ESA it might be possible to use spectroscopy to identify atmospheric signatures suggesting plate tectonism on these objects. This class of planets offers the possibility of finding Earth analogs and, in particular, make attractive targets in the search for habitable planets." }, "0710/0710.2655_arXiv.txt": { "abstract": "The detector material Cadmium Zinc Telluride (CZT) achieves excellent spatial resolution and good energy resolution over a broad energy range, several keV up to some MeV. Presently, there are two main methods to grow CZT crystals, the Modified High-Pressure Bridgman (MHB) and the High-Pressure Bridgman (HPB) process. The study presented in this paper is based on MHB CZT substrates from the company Orbotech Medical Solutions Ltd. \\cite{Orbotech}. Former studies have shown that high-work-function materials on the cathode side reduce the leakage current and therefore improve the energy resolution at lower energies. None of the studies have emphasized on the anode contact material. Therefore, we present in this paper the result of a detailed study in which for the first time the cathode material was kept constant and the anode material was varied. We used four different anode materials : Indium, Titanium, Chromium and Gold, metals with work-functions between 4.1~eV and 5.1~eV. The detector size was 2.0$\\times$2.0$\\times$0.5~cm$^3$ with 8$\\times$8 pixels and a pitch of 2.46~mm. The best performance was achieved with the low work-function materials Indium and Titanium with energy resolutions of 2.0~keV (at 59~keV) and 1.9~keV (at 122~keV) for Titanium and 2.1~keV (at 59~keV) and 2.9~keV (at 122~keV) for Indium. Taking into account the large pixel pitch of 2.46~mm, these resolutions are very competitive in comparison to those achieved with detectors made of material produced with the more expensive conventional HPB method. We present a detailed comparison of our detector response with 3-D simulations. The latter comparisons allow us to determine the mobility-lifetime-products ($\\mu\\tau$-products) for electrons and holes. Finally, we evaluated the temperature dependency of the detector performance and $\\mu\\tau$-products. For many applications temperature dependence is important, therefore, we extended the scope of our study to temperatures as low as -30$^{\\circ}$C. There are two important results. The breakdown voltage increases with decreasing temperature, and electron mobility-life-time-product decreases by about 30\\% over a range from 20$^{\\circ}$C to -30$^{\\circ}$C. The latter effect causes the energy resolution to deteriorate, but the concomitantly increasing breakdown voltage makes it possible to increase the applied bias voltage and restore the full performance. ", "introduction": "\\label{Introduction} Cadmium Zinc Telluride (CZT) has emerged as the material of choice for the detection of hard X-rays and soft gamma-rays with excellent position and energy resolution and without the need for cryogenic cooling. The high density of CZT ($\\rho\\,\\simeq$~5.76 g/cm$^3$) and high average atomic number ($\\simeq$~50) result in high stopping power and in a large cross section for photoelectric interactions. The main application of CZT detectors is the detection of photons in the 10~keV to $\\sim$1~MeV energy range. CZT has a major impact in various fields including medical imaging, homeland security applications, and space-borne X-ray and gamma-ray astronomy. To give a few examples of the latter, the Swift mission launched in 2004 carries a wide field of view X-ray telescope for the discovery of gamma-ray bursts (GRBs) in the energy range from 15~keV to 150~keV \\cite{swift}. This Burst Alert Telescope makes use of the coded mask technique to localize GRBs with an accuracy of 1-2~arcmin. It is built out of an array of 32,768 co-planar 2$\\times$4$\\times$4~mm$^3$ CZT detectors covering a total area of 0.5~m$^2$. The proposed EXIST (Energetic X-ray Imaging Survey Telescope) mission \\cite{Grindlay} also uses the coded mask approach. With 15,000 pixelated CZT detectors, each with a volume of 2.0$\\times$2.0$\\times$0.5~cm$^3$ and with 16$\\times$16 pixels, an angular resolution of 5 arcmin will be achieved and the energy range between 10~keV and 600~keV will be covered.\\\\ In this paper, we present a detailed study of CZT grown with the Modified Horizontal Bridgman (MHB) process by the company Orbotech Medical Solutions Ltd. \\cite{Orbotech} (previously Imarad). This growth technique gives uniform substrates at high yields and thus modest costs. One of the disadvantages of the MHB technique is a somewhat lower bulk resistivity of $10^9\\,\\Omega\\,$cm compared to $10^{10}\\,\\Omega\\,$cm obtained with the more conventional High Pressure Bridgman (HPB) process. This results in higher leakage currents and therefore degrades the energy resolution at lower energies ($<$ 100~keV). Several authors recognized that it is possible to reduce dark currents in MHB detectors by using {\\sl p}-type-intrinsic-{\\sl n}-type (PIN) and metal-semiconductor-metal (MSM) contacts \\cite{Nemirovski:01,Vadawale:04,Nari:98,Nari:00,Nari:02}. Both schemes, PIN and MSM, can indeed reduce the dark currents by factors $>$10 and improve on the energy resolution of the detectors at low ($<$ 100 keV) energies. Encouraged by the result of these authors, we experimented with different cathode and anode materials on several MHB substrates, optimizing for the first time the cathode and anode contacts separately. We discussed a comparison of different cathode materials in an earlier paper \\cite{Jung:05}. Gold (Au) was among the metals yielding the best performance in terms of dark current suppression and energy resolution as cathode contact material. With the objective to finalize our studies on contact materials, we present in this paper the results on different anode contact materials with an Au cathode and added a study of the performance at lower temperatures which is especially important for space borne applications. We derived $\\mu\\tau$-products of electrons and holes needed for detector simulations. These reproduce the measured data well, and can be used to further improve detector performance and design. The paper is structured as follows: \\\\ In section \\ref{sec:detectorfabrication}, origin and some general properties of our substrate are presented. We explain how we used this substrate to fabricate detectors suited for our measurements and give a short description of our experimental setup. Then in section \\ref{sec:char}, we discuss crystal inhomogeneities and present the results of photoluminescence mapping of the Zn content and of infrared transmission microscopy. Subsequently, in section \\ref{sec:contact}, the choice of the contact materials is explained, and the design of the detectors we used for comparison measurements is described in detail. Section \\ref{sec:performance} presents the results of our comparison experiments and a discussion of the influence of anode contact materials on the performance. In section \\ref{sec:Simulation}, we show that we can reproduce experimental data with computer simulations. For this, $\\mu\\tau$-products of electrons and of holes need to be known. We explain how we measured the $\\mu\\tau$-products of electrons and used simulations to derive the $\\mu\\tau$-products of holes. In section \\ref{sec:T}, we present studies of the temperature dependence on the detector response and the electronic properties of CZT. ", "conclusions": "\\label{sec:disc} In this paper, we present the results of a detailed study of the performance of a MHB CZT detector from the company Orbotech Medical Solutions Ltd. \\cite{Orbotech}. \\\\ Using photoluminescence, we mapped the spatial distribution of the zinc content, which showed a variation of $\\sim$3\\%. This variation does not affect the energy resolution of the detector if information about the location of the interaction is available (e.g. pixelated detectors), because in this case the effect can be corrected for. With infrared transmission microscopy, we could detect some crystal defects and could correlate poor pixel performance with such defects. \\\\ We tested various anode contact materials and found that Ti showed the best performance of 2.0~keV, 1.9~keV and 7.3~keV at 59~keV, 122~keV and 662~keV, respectively. We are using this metal now to contact all our MHB detectors. We showed detailed comparisons of experimentally measured and simulated detector response. After adjusting the hole mobility and lifetime, our simulation gives a good description of the measured data and shows that we have a deep understanding of our data. Therefore, the simulation can be used to optimize the detector design as the pixel width and pitch.\\\\ We studied the detector properties as a function of temperature. The electron $\\mu\\,\\tau$-product drops by about 30\\%, when cooling the substrate from 20$^{\\circ}$C to -30$^{\\circ}$C with constant bias voltage, and the energy resolution deteriorates accordingly, a behavior that has been noted already for MHB substrates \\cite{Nari:00}. For HPB substrates, a corresponding decrease of the electron $\\mu\\,\\tau$-product has been observed \\cite{Stur:05}. We were able to recover almost all of the room temperature performance by increasing the bias voltage at low temperatures, which is possible, because at lower temperatures the breakdown voltage is increased. Taking everything into account, we conclude, that the overall performance of CZT detectors, produced with the cost-effective MHB process, is comparable to the performance of CZT detectors which are produced with the expensive HPB process. \\\\[2ex] {\\it Acknowledgments:} We thank Uri El Hanany from Orbotech Inc. for several free CZT detectors. We acknowledge S. Komarov, L.~Sobotka, D.~Leopold, and J.~Buckley for helpful discussions. Thanks to electrical engineer P.~Dowkontt, and electrical technician G.~Simburger for their support. This work is supported by NASA under contracts NNG04WC176 and NNG04GD70G, and the NSF/HRD grant no.\\ 0420516 (CREST). The authors at Fisk University gratefully acknowledge financial support from the National Science Foundation through the Fisk University Center for Physics and Chemistry of Materials (CPCoM), Cooperative Agreement CA: HRD-0420516 (CREST program), and from US DOE through the National Nuclear Security Administration (NNSA), Office of Nonproliferation Research and Engineering (NA-22), grant no. DE-FG52-05NA27035." }, "0710/0710.4186_arXiv.txt": { "abstract": "{ Existence of GZK neutrinos (ultra high energy neutrinos) have been justified although the flux is very low. % A new method is desired to use a huge mass of a detector medium to detect them. % A fundamental study of radar method was carried out to measure % microwave reflection from electromagnetic energy deposit by X-ray irradiation % in a small rock salt sample. % The reflection rate of $1 \\times 10^{-6}$ was found at the energy deposit of $ 1\\times 10^{19}$ eV % which was proportional to square of the X-ray intensity suggesting the effect to be coherent scattering. % The decay time of the reflection was several seconds. This effect implies a large scale natural rock salt formation could be utilized like % a bubble chamber irradiated by radio wave instead of visible light to detect GZK neutrinos. \\PACS{ {61.80.Cb}{X-ray effects} \\and {95.55Vj}{Neutrino, pion, and other elementary particle detectors: cosmic ray detectors} } % } % ", "introduction": "\\label{intro} GZK (Greisen, Zatsepin and Kuzmin) neutrinos are generated by collision between % ultra high energy (UHE) cosmic rays ($\\geq 4 \\times 10^{19}$ eV) and cosmic microwave background % Ref.~\\cite{Greisen}. % Both have been observed, then GZK neutrinos ($\\geq 10^{16}$ eV) % have been justified to exist although the flux is very low ($\\sim$ 1 km$^{-2}$ day$^{-1}$). % The flux can be estimated to a certain extent % since flux of UHE cosmic rays and density of cosmic microwave background are known. % GZK neutrinos could become a standard candle of UHE neutrinos for finding out other % conceivable UHE neutrino sources such as active galactic nuclei, topological defects and $\\gamma$-ray bursts etc. % in the universe. % In order to detect GZK neutrinos, a huge mass of a detection medium as large as % 50 Gt (3 $\\times$ 3 $\\times$ 3 km$^3$ for a rock salt case) is needed due to their low flux. % Detecting radio wave from Askryan effect (coherent Cherenkov effect) Ref.~\\cite{Askaryan} is a % promising way to utilize a large mass of natural rock salt Ref.~\\cite{Stanley} or % ice bed in Antarctica % but it needs a lot of bore holes to be installed for radio wave detection antennas. We had measured attenuation length of natural rock salt samples as well as % synthetic rock salt samples of single crystal by a perturbed cavity resonator method. % We found long attenuation lengths for electric field % more than 200 m at 0.3 - 1 GHz in the samples of natural rock salt % Ref.~\\cite{Chiba} as shown in fig. \\ref{fig:attenuation}. % Data noted as \"Hockley\\_in\\_situ\" in the legend are % in-situ measurements at Hockley salt mine Ref.~\\cite{Hockley}. The long attenuation length allows us to % utilize a rock salt formation as a UHE neutrino detector Ref.~\\cite{Gorham}. \\begin{figure*} \\includegraphics[width=1.\\textwidth,height=0.68\\textwidth,angle=0]{attenuation.eps} \\caption{Attenuation lengths for electric field are plotted with respect to frequency for % synthetic and natural rock salt. % The upper and the lower straight lines are % fitted to synthetic and natural rock salt samples of Asse salt mine, respectively, % assuming loss tangent is constant with respect to the frequency.} \\label{fig:attenuation} % \\end{figure*} A basic study of a new detection method has been carried out to measure % radio wave reflection from electromagnetic energy deposit by X-ray irradiation % in a small rock salt sample. % If we could detect the radio wave reflection by an enough reflection rate, a large scale natural rock salt formation would work as % a radio bubble chamber using radio wave instead of visible light to get % 3 dimensional information of the energy deposit of the shower generated by the UHE neutrino. ", "conclusions": "\\label{sec:3} Microwave was reflected at the rate of $10^{-6}$ from X ray irradiated rock salt % in the energy deposit of $1\\times 10^{19}$ eV. % Life time of particles which was responsible to the decay time of % the microwave reflection was several seconds. % It is long enough to employ periodic transmission of radar pulses without triggered by % reception of Askaryan radio wave. % Microwave scattering from the irradiated rock salt was coherent % but the particle species working as the scattering targets is not known. % Power of radio wave emitted by Askaryan effect % increases proportionally to the frequency. % At the higher frequency the attenuation length becomes short % as shown in fig. \\ref{fig:attenuation}. % On the contrary radar method is not imposed such a restriction on the frequency. % Strong artificial radar pulses are available for the radar method. % Consequently, we could get long range of detection. % Times and amplitude from several receiving antennas could give us 3 dimensional % information of the UHE shower with its energy. % In order to utilize a salt dome with a diameter of 3 km and a depth of 3 km (50 Gt), % several bore holes are needed in which the transmitting and receiving antennas are installed. Radar method would have a potential to realize Salt Neutrino Detector to act like a Radio % Bubble Chamber to detect GZK neutrinos. % If we could get the peak power of the radar considerably larger than 1 GW, % the range becomes long enough to know whether GZK neutrinos exist or not without expensive bore holes. % The antennas would be installed slightly under a floor in an excavated space of a rock salt dome. % \\begin{acknowledgement} Work is partially supported by a Grant in Aid for Scientific Research for Ministry of Education, Science, % Technology and Sports and Culture of Japan, and Funds of Tokubetsu Kenkyuhi, at Seikei University. % We express deep appreciation to KEK-PF staffs, especially Dr. Kazuyuki Hyodo who extended us his hospitality % throughout this experiment, without him this measurement could not be carried out smoothly. \\end{acknowledgement}" }, "0710/0710.4465_arXiv.txt": { "abstract": "We study the effect of the neutron star spin -- kick velocity alignment observed in young radio pulsars on the coalescence rate of binary neutron stars. Two scenarios of the neutron star formation are considered: when the kick is always present and when it is small or absent if a neutron star is formed in a binary system due to electron-capture degenerate core collapse. The effect is shown to be especially strong for large kick amplitudes and tight alignments, reducing the expected galactic rate of binary neutron star coalescences compared to calculations with randomly directed kicks. The spin-kick correlation also leads to a much narrower NS spin-orbit misalignment. ", "introduction": "There is an increasing interest of a broad astrophysical community to coalescing binary compact stars as primary sources of gravitational waves for the ground-based gravitational wave observatories. Double neutron stars (DNS), observed as binary pulsars, remain to be the most reliable objects for gravitational wave searches. On-going LIGO science runs \\citep{Abbot07} have already set first experimental upper limits on their galactic rates of a few per year. Astrophysical estimates of the DNS coalescence rate, which are based on the binary pulsar statistics or can be obtained from population synthesis simulations, are model-dependent and vary within more than an order of magnitude around the value $10^{-5}$ per year (see recent reviews \\citealt{PYu,Kalogera&07} and references therein). The kick velocity imparted to a newborn neutron star is an important phenomenological parameter of the core collapse supernovae and represents one of the major uncertainties in the theory of binary star evolution. The origin of the kicks remains unclear and a number of physical models have been suggested (see, for example, \\citealt{Lai} and references therein). For post-supernova evolution of a binary, both the amplitude of the kick and its space direction are important. The distribution of the kick amplitudes is usually obtained from the analysis of radio pulsar proper motions \\citep{Hobbs}. The direction of kicks (for example, with respect to the spin axis of the neutron star) is more difficult to infer from observations. Recently, several observational clues appeared indicating possible NS spin-kick alignment. A noticeable spin-kick alignment has been inferred from polarization measurements of radio emission of pulsars \\citep{Johnston_ea, Rankin07, Johnston_ea07}, as well as from X-ray observations of pulsar wind nebulae around young pulsars \\citep{Helfand,Kargaltsev}. Implications of these findings to the formation of double pulsars were discussed by \\cite{Wang_ea06}. The possibility and conditions for such an alignment in the model of the kick origin by multiple random kicks during NS formation (proposed by \\citep{Spruit}) were studied by \\cite{Wang_ea07}. The implication of NS kick-spin correlation to the plausible birth-kick scenarios was also discussed by \\cite{Ng&Romani}. Here we explore the effect of NS spin -- kick correlation on the formation and galactic coalescence rate of double neutron stars (DNS) which are primary targets for modern gravitational wave detectors. We show that the tighter alignment, the smaller is the DNS merging rate with respect to models with random kick orientation. The effect is especially important for large kick amplitudes ($\\sim 400$ km/s). We calculate the spin-orbit misalignment of the components of DNS which can be important for GW data analysis. We also considered a scenario in which no (or insignificant) kick accompanies the formation of a neutron star in binary systems from the main-sequence progenitors in a restricted mass range (8-11 $M_\\odot$ or so) due to electron-capture collapse of O-Ne-Ng degenerate stellar core proposed by \\cite{Podsiadlowski_ea04} and further elaborated by \\cite{vdH04,vdH07}. This hypothesis is phenomenologically based on the existence of long-period Be X-ray binaries with low eccentricities \\citep{Pfahl_ea02}. It is consistent with the evolutionary analysis of double neutron star formation \\citep{vdH07} and has been used in some population synthesis studies of DNS, see for example \\cite{Dewi_ea05, Dewi_ea06}. ", "conclusions": "We have shown that the spin-velocity correlation observed in radio pulsars, suggesting the NS spin-kick velocity alignment, may have important implications to GW studies. First, the tight alignment reduces the galactic rate of double neutron star coalescences (especially for large kicks 300-400 km/s) relative to models with random kicks. Second, the spin-kick correlation results in a specific distribution of NS spin -- orbit misalignments. In turn, analysis of the NS spin-orbit misalignments inferred from GW signals during DNS mergings can be potentially used to put independent bounds on the still elusive nature of NS kicks." }, "0710/0710.3590_arXiv.txt": { "abstract": "We review recent results of SPH simulations of gravitational instability in gaseous protoplanetary disks, emphasizing the role of thermodynamics in both isolated and binary systems. Contradictory results appeared in the literature regarding disk fragmentation at tens of AU from the central star are likely due to the different treatment of radiation physics as well as reflecting different initial conditions. Further progress on the subject requires extensive comparisons between different codes with the requirement that the same initial conditions are adopted. It is discussed how the local conditions of the disks undergoing fragmentation at $R < 25$ AU in recent SPH simulations are in rough agreement with the prediction of analytical models, with small differences being likely related to the inability of analytical models to account for the dynamics and thermodynamics of three-dimensional spiral shocks. We report that radically different adaptive hydrodynamical codes, SPH and adaptive mesh refinement (AMR), yield very similar results on disk fragmentation at comparable resolution in the simple case of an isothermal equation of state. A high number of refinements in AMR codes is necessary but not sufficient to correctly follow fragmentation, rather an initial resolution of the grid high enough to capture the wavelength of the strongest spiral modes when they are still barely nonlinear is essential. These tests represent a useful benchmark and a starting point for a forthcoming code comparison with realistic radiation physics. ", "introduction": "Physical fragmentation in astrophysical systems such as gravitationally unstable protoplanetary disks depends on the competition between gravity and thermal pressure, at least in the limit in which the contribution of magnetic fields is neglected. Knowing whether existing numerical simulations are modeling correctly both gravity and pressure is then of paramount importance in this context. Thermal pressure is affected by the details of cooling and heating, either by radiation, convection or viscosity. The increasingly more sophisticated simulations designed in the last few years have explored the effect of thermodynamics on disk fragmentation. finding in general that this is less likely than in older models in which the gas was evolved using a locally isothermal equation of state (Boss 2002a,b; Mayer et al. 2002; Rice et al. 2003). Currently, it is debated whether fragmentation into Jupiter-sized clumps is possible, especially at distances less than $50$ AU from the central star (Durisen et al. 2007; Stamatellos \\& Whitworth 2007). It is well understood that the disk must cool on a timescale comparable to the orbital time in order for fragmentation to occur (Rice et al. 2003, Gammie 2001; Johnson \\& Gammi 2003; Clarke et al. 2007). At a few tens of AU the optically thick disk midplane cools via radiation on a timescale longer than the local orbital time, hence the only chance for disks to cool efficiently is via a non-radiative mechanism. This could be either convection or turbulent diffusion associated with shock bores (Boss 2004; Mayer et al. 2007; Boley \\& Durisen 2006). Finally, analytic calculations that include convection predict no fragmentation at radii $< 50$ AU for disks with masses significantly smaller than the mass of the central star, as it is the case in T Tauri disks (Rafikov 2005, 2007). Four groups have implemented a scheme for radiative transfer in three dimensional simulations. Two of them, using, respectively, an SPH and a finite-difference polar grid code, and similar initial conditions, find that fragmentation can happen at $R < 20$ AU (Boss 2004; Mayer et al. 2006), while the two other groups use , respectively, an SPH (Stamatellos \\& Whitworth 2007) and a cylindrical grid code (Boley et al. 2006) with nearly identical initial conditions and find that disk fragmentation does not happen at $R < 50$ AU (Stamatellos \\& Whitworth 2007 find that fragmentation is possible at larger radii, $R \\sim 100$ AU). These contradictory results were obtained with four different codes. The two sets of initial conditions were also significantly different, the disks in Mayer et al. (2007) and Boss (2006) having much higher surface densities than those in Boley et al. (2006) and Stramatellos \\& Whitworth (2007), and thus being more prone to fragmentation. Numerical codes differ not only in the way they implement radiative transfer but also in more basic aspects of the algorithm such as how they compute gravity and how they solve the energy equation. Further progress in elucidating whether disk instability is a possible formation mechanism for giant planets clearly require that the different codes adopted in this area are compared on identical conditions, first in simple models with a fixed equation of state and then on progressively more sophisticated models with radiative transfer. Here we summarize the results recently obtained with the SPH code GASOLINE (Wadsley, Stadel \\& Quinn 2004) with a scheme for radiative transfer for the case of both isolated and binary systems. Then we discuss the results of the first stage of a code comparison project in which the isothermal disks are evolved with GASOLINE and with one of the most advanced grid codes currently available, the adaptive mesh refinement (AMR) code FLASH (Fryxell et al. 2000). ", "conclusions": "" }, "0710/0710.0300_arXiv.txt": { "abstract": "{In the last couple of years a population of very massive ($M_\\star>10^{11}$ M$_\\odot$), high-redshift ($z\\ge2$) galaxies has been identified, but its role in galaxy evolution has not yet been fully understood.} {It is necessary to perform a systematic study of high-redshift massive galaxies, in order to determine the shape of the very massive tail of the stellar mass function and determine the epoch of their assembly.} {We selected high-$z$ massive galaxies at 5.8$\\mu$m, in the SWIRE ELAIS-S1 field (1 deg$^2$). Galaxies with the 1.6$\\mu$m stellar peak redshifted into the IRAC bands ($z\\simeq1-3$, called ``IR-peakers'') were identified. Stellar masses were derived by means of spectro-photometric fitting and used to compute the stellar mass function (MF) at $z=1-2$ and $2-3$. A parametric fit to the MF was performed, based on a Bayesian formalism, and the stellar mass density of massive galaxies above $z=2$ determined.} {We present the first systematic study of the very-massive tail of the galaxy stellar mass function at high redshift. A total of 326 sources were selected. The majority of these galaxies have stellar masses in excess of $10^{11}$ M$_\\odot$ and lie at $z>1.5$. The availability of mid-IR data turned out to be a valuable tool to constrain the contribution of young stars to galaxy SEDs, and thus their $M_\\star/L$ ratio. The influence of near-IR data and of the chosen stellar library on the SED fitting are also discussed. The $z=2-3$ stellar mass function between $10^{11}$ and $\\sim10^{12}$ M$_\\odot$ is probed with unprecedented detail. A significant evolution is found not only for galaxies with $M\\sim10^{11}$ M$_\\odot$, but also in the highest mass bins considered. The comoving number density of these galaxies was lower by more than a factor of 10 at $z=2-3$, with respect to the local estimate. SWIRE 5.8$\\mu$m peakers more massive than $1.6 \\times10^{11}$ M$_\\odot$ provide 30$-$50\\% of the total stellar mass density in galaxies at $z=2-3$. } {} ", "introduction": "\\label{sect:intro} Tracing the formation of galaxies and understanding the epoch {\\em when} the bulk of their baryonic mass was assembled represents one of the major problems of modern cosmology, particularly controversial when dealing with massive ($M_{\\textrm{stars}} > 10^{11}$ M$_\\odot$) objects. The assembly of massive galaxies is one of the critical questions in the cosmic evolutionary scenario. The uniform properties of local early-type galaxies and of the fundamental plane have inspired the so called ``monolithic collapse'' scenario \\citep{eggen1962,chiosi2002}, in which galaxies formed in the remote past through huge events of star formation and subsequently evolved passively across cosmic time. On the other hand, in the more recent ``hierarchical'' scenario \\citep{white1978,kauffmann1996,kauffmann1998,somerville1999}, massive galaxies assemble by mergers of lower-mass units, with the most massive objects being born in the latest stages of evolution, at $z\\le1$. The availability of several powerful tracers of star formation (e.g. UV continuum, optical recombination lines, far-IR emission, sub-mm light) has favored the popularity of studies of the comoving star formation density \\citep{madau1996,lilly1996} in galaxies at various redshifts. It is now well determined that the Universe experienced an epoch of enhanced star formation in the past, peaking at $z\\simeq1-2$, with a subsequent decline of at least one order of magnitude to the present time \\citep[e.g.][ among others]{hopkins2004,rudnick2003,flores1999,madau1996}. An alternative approach consists of studying the mass already assembled in galaxies, instead of the amount of stars being formed. The integral of the past star formation density provides the stellar mass density at a given epoch: a complementary constraint on cosmic galaxy evolution. Thus, the build up of the stellar mass across cosmic time has become one of the major topics in observational cosmology, and has overtaken the classic Madau-Lilly diagram as the central tool for studying galaxy evolution. The large observational effort dedicated to this subject has shown that the global stellar mass density increases from early epochs to the low-redshift Universe \\citep[e.g.][]{brinchmann2000,dickinson2003,fontana2003, fontana2004,fontana2006,rudnick2003,rudnick2006,drory2005}. Very deep surveys have been exploited to describe the shape of the stellar mass function at high redshift \\citep{fontana2006,drory2005,gwyn2005}, but a clear picture on the role of very massive galaxies has not yet emerged. Several pieces of evidence exist that fully formed massive galaxies were already in place at redshift $z\\sim 2-3$. A substantial population of luminous red galaxies at redshifts $z>2$ (known as ``distant red galaxies'', DRGs) was found in the Faint InfraRed Extragalactic Survey \\citep[FIRES,][]{franx2003}. Based on near-IR spectroscopy \\citep{foerster2004,vandokkum2004}, these galaxies turned out to be massive ($M=1-5\\times10^{11}$ M$_\\odot$), evolved (ages of $1-2.5$ Gyr) systems, probably descendants of galaxies which started forming at redshift $z>4$. Based on FIRES data, \\citet{rudnick2003} inferred that DRGs contribute $\\sim50$\\% of the global stellar mass density at $z=2-3$. \\citet{dickinson2003} exploited Hubble Deep Field North NICMOS data to derive the stellar masses of galaxies up to $z=3$. The study of the global stellar mass density highlighted that $50-75$\\% of the present-day stellar mass was already in place at $z\\simeq1$, while only $3-14$\\% had been already assembled at $z\\simeq2.7$. Direct determinations of the galaxy mass function based on near-IR deep imaging by the Spitzer Space Telescope indicate that the number of massive galaxies does not significantly evolve up to at least $z\\simeq1$ \\citep{franceschini2006,fontana2006,bundy2005}. Using data from the Gemini Deep Deep Survey \\citep[GDDS,][]{abraham2004}, \\citet{glazebrook2004} identified a population of red ($[I-K]>4$ Vega mag.) galaxies at $z\\simeq2$ with stellar masses in excess of $10^{11}$ M$_\\odot$. These objects contribute roughly 30\\% of the total stellar mass density of the Universe at that epoch. \\citet{mccarthy2004} estimated the age of red galaxies at $z=1.3-2.2$ in the GDDS, deriving a median age of 1$-$3 Gyr, and a star formation history dominated by very powerful bursts (300$-$500 M$_\\odot$ yr$^{-1}$). These massive galaxies must have undergone a rapid formation process at $z>1$. Exploiting data from the K20 survey \\citep{cimatti2002c,cimatti2002b,cimatti2002a}, \\citet{daddi2004} identified few luminous $K$-band selected galaxies at $1.7 < z < 2.3$ with stellar masses $M \\simeq 10^{11}-5\\times 10^{11}\\ [M_\\odot]$. Combining deep K20 spectroscopy and HST-ACS imaging, \\citet{cimatti2004} discovered four old, fully assembled spheroidal galaxies at $1.6 < z < 1.9$: the most distant such objects currently known. The stellar mass of these galaxies turned out to be in the range $1-3\\times10^{11}$ M$_\\odot$. \\citet{fontana2004} studied K20 galaxies as well, showing that massive ($M>10^{11}$ M$_\\odot$) galaxies are easily found up to $z\\simeq2$. These authors also report on the stellar mass function: only mild evolution ($\\sim$2$-$30\\%) is detected to $z=1$, but only $\\sim35$\\% of the $z=0$ stellar mass locked up in massive objects was assembled by $z=2$. At even higher redshifts, \\citet{rigopoulou2006} have recently studied a population of $z\\sim3$ Lyman-break galaxies (LBGs) with stellar masses in excess of $10^{11}$ M$_\\odot$. \\citet{mclure2006} identified nine LBGs at $z\\ge5$ in the UKIDSS \\citep{lawrence2006} survey, over an area of 0.6 deg$^2$. A stacking analysis suggests that the typical stellar mass of these sources is $>5 \\times 10^{10}$ M$_\\odot$. \\citet{mobasher2005} analyzed the properties of J-dropouts in the Hubble Ultra Deep Field \\citep[HUDF,][]{beckwith2006}, exploiting Spitzer photometry. They identified a $z\\sim6.5$ candidate that was interpreted as a post-starburst galaxy with a surprisingly high stellar mass of $5.7 \\times 10^{11}$ M$_\\odot$ (but see, for example, Yan et al. \\citeyear{yan2004} for a different interpretation). \\citet{drory2005} studied the stellar mass function of galaxies in the FORS Deep Field \\citep[FDF,][]{heidt2003} and GOODS/CDFS \\citep{giavalisco2004} field, over a total area of 90 arcmin$^2$ and found that the total stellar mass density at $z=1$ is 50\\% of the local value. At $z=2$, 25\\% of the local mass density was already assembled, and at $z=3$ and $z=5$, at least 15\\% and 5\\% of stellar mass, respectively, was already in place. Massive ($M>10^{11}$ M$_\\odot$) galaxies existed over the whole redshift range probed, up to $z=5$. The number density of these massive galaxies evolves very similarly to galaxies with $M>10^{10}$ M$_\\odot$, decreasing by 0.4 dex to $z=1$, 0.6 dex to $z=2$, and 1 dex to $z=4$. By analyzing the properties of $K$-selected galaxies in the $\\sim131$ arcmin$^2$ of the GOODS-CDFS survey, \\citet{caputi2006} found that the vast majority ($85-90$\\%) of local $M > 2.5 \\times 10^{11}$ M$_\\odot$ galaxies appears to be already in place at $z\\sim1$. These authors also infer that roughly $65-70$\\% of these galaxies assembled at $z=1-3$ by means of obscured, intense bursts of star formation, while the remaining could be in place at even higher redshifts ($z= 3-4$). The observational challenge is that large volumes are needed to find representative samples of such rare very massive galaxies at high redshift. The Spitzer Wide-area InfraRed Extragalactic survey \\citep[SWIRE,][]{lonsdale2003,lonsdale2004} observed $\\sim$49 deg$^2$ in the seven Spitzer channels, and is therefore ideal to find rare objects. Its volume is large enough to detect $\\sim$85 DM haloes of mass $> 10^{14} M_\\odot$ in the $210^{12}\\ L_\\odot$) and most massive ($M_\\star>\\textrm{several}\\ 10^{11}\\ M_\\odot$) galaxies ever to exist. The Infrared Array Camera \\citep[IRAC,][]{fazio2004}, onboard Spitzer \\citep{werner2004}, samples the restframe near-IR light of distant galaxies. A near-IR selection not only directly probes the low-mass stars dominating the baryonic mass of a galaxy, but also is minimally affected by dust extinction. Therefore Spitzer is best suited to the study of the stellar content of galaxies up to $z=3$. Moreover Spitzer allows detection of galaxies that would be missed by restframe UV selection, for example distant red galaxies \\citep[e.g.][]{daddi2004}. We take advantage of the shape of near-IR spectral energy distribution (SEDs) of galaxies to identify high-redshift objects on the basis of IRAC colors. Our selection is based on the detection of the 1.6$\\mu$m stellar peak in galaxies \\citep{sawicki2002,simpson1999}, redshifted to the IRAC domain. It is also worth noting that galaxies selected in this way benefit from a negative k-correction in the IRAC bands, because the slope of a galaxy's SED is negative redward of the 1.6$\\mu$m peak. In this way we performed a systematic search for $M\\gtrsim10^{11}$ M$_\\odot$ galaxies at $z>1$. The analysis is carried out in the central square degree of the ELAIS-S1 SWIRE field, where optical, near-IR ($J$, $K_s$) photometry and optical spectroscopy are available (Berta et al., \\citeyear{berta2006}, Dias et al., in prep., La Franca et al., in prep.). This area, and the sampled volume, are bigger than any other previously explored for studying very massive galaxies at $z=1-3$, which were limited to very deep, pencil-beam surveys \\citep[e.g.][]{dickinson2003,drory2005,gwyn2005,fontana2006}. This paper is structured as follows. Section {\\bf 2} presents the data available in the ELAIS-S1 field; in Sect. {\\bf 3} we present our selection criterion; then Sect. {\\bf 4} discusses the photometric estimate of redshifts. Section {\\bf 5} deals with the estimate of the stellar mass in galaxies, and presents a very detailed analysis of the influence of mid-IR and near-IR constraints on it. Experiments with different stellar libraries and IMFs are also discussed. Section {\\bf 6} presents the stellar mass function of our galaxies, including completeness correction and a parametric fit based on a Bayesian formalism. Finally, Sects. {\\bf 7} and {\\bf 8} discuss results and draw our conclusions. Throughout this work, we adopt a standard $H_0=71$ $[$km s$^{-1}$ Mpc$^{-1}]$, $\\Omega_m=0.27$, $\\Omega_\\Lambda=0.73$ cosmology, unless otherwise stated. ", "conclusions": "Sampling the very massive tail of the stellar mass function at high redshift and estimating its contribution to the global stellar mass density is a critical task in modern cosmology, motivated by the recent evidence that a ``downsizing'' effect exists in the evolution of stellar mass across cosmic time. We have exploited SWIRE/Spitzer and ancillary data in the ELAIS-S1 area to perform a systematic search for high redshift ($z\\gtrsim1.0$) massive ($M>10^{11}$ M$_\\odot$) galaxies. High redshift systems have been isolated by identifying the 1.6$\\mu$m restframe stellar peak shifted to IRAC wavelengths (3.6$-$8.0 $\\mu$m). The availability of near-IR ($J$ and $K_s$ band) data allowed us to avoid low-redshift interlopers and selection aliasing due to bright 3.3 $\\mu$m PAH features at $z\\simeq0.4$. A total of 203 5.8$\\mu$m-peakers and 123 4.5$\\mu$m-peakers have been identified over one square degree in the ELAIS-S1 field. We have performed an extensive SED analysis, based on mixed stellar population synthesis, focused on deriving the stellar masses of the selected sample. The advantage of near-IR and mid-IR constraints, as well as the dependence of results on the choice of the IMF and SSP library have been explored in detail. The main results are: \\begin{itemize} \\item because of the shallow 5.8$\\mu$m flux cut adopted, the SWIRE IR-peaker sample consists of very massive galaxies, the majority of sources having $M_\\star>10^{11}$ M$_\\odot$. Objects in the range $z=1-2$ peak at $M_\\star\\simeq10^{11}$ M$_\\odot$, while the distribution of 5.8$\\mu$m-peakers ($z=2-3$) is centered at $M_\\star\\simeq2\\times 10^{11}$ M$_\\odot$. Typical uncertainties in stellar mass estimate (due to degeneracies in the SFH space) range between 0.1 and 0.3 dex, depending on multiwavelength coverage. The emission of $\\sim$30\\% of the sources turns out to be dominated by stars older than 1 Gyr, and also in the majority of the remaining cases old stellar populations do contribute to the observed SEDs. \\item the availability of mid-IR data provides a valuable constraint on the recent star formation history of individual galaxies. If no 24$\\mu$m flux (nor upper limit) were available, the resulting stellar masses could be underestimated and the spread in mass would be wider. \\item despite being very useful in the selection process, near-IR ($J$ and $K_s$) data turned out to be effective in constraining the D4000 break only in $\\sim$15\\% of cases, and preferentially for 4.5$\\mu$m-peakers. \\item the choice of a \\citet{chabrier2003} IMF, instead of a \\citet{salpeter1955} one, leads to systematically lower stellar masses, resulting in a rigid shift of the stellar mass distribution by $\\sim0.3$ dex. \\item when including thermally-pulsing AGB stars in the SSP library \\citep{maraston2005}, the fraction of objects dominated by old stellar populations increases, but the overall stellar mass distribution does not change significantly, because of the different $M_\\star/L$ ratios of SSPs. \\end{itemize} The stellar mass estimates have been used to compute the comoving number density of galaxies as a function of stellar mass (i.e. the observed stellar mass function), adopting the accessible volume formalism. Following \\citet{fontana2004}, we have corrected the samples for mass incompleteness and recovered the mass function down to $1.25\\times10^{11}$ and $1.6\\times10^{11}$ M$_\\odot$ for 4.5$\\mu$m- and 5.8$\\mu$m-peakers respectively. Unfortunately, the selection, based on a 5.8$\\mu$m flux cut turned out to be only partially sensitive to 4.5$\\mu$m peakers, therefore only a lower limit on the mass function of $z=1-2$ sub-sample could be set. The observed stellar mass function of 5.8$\\mu$m-peakers was reproduced with a parametric function, using the STY \\citep{sandage1979} approach. The uncertainties in the stellar masses of individual sources were automatically included in the analysis by using a Bayesian formalism, and the best fit was obtained with a MCMC sampling of the parameter space. The stellar mass function was finally integrated to derive the stellar mass density locked in 5.8$\\mu$m-peakers with $M>1.6\\times10^{11}$ M$_\\odot$, at $z=2-3$. The results of the mass function analysis are: \\begin{itemize} \\item the wide area surveyed by SWIRE allows the very massive tail of the stellar mass function to be probed with unprecedented detail at $z=2-3$, extending previous analyses up to $M_\\star=7\\times10^{11}$ $[$h$_{70}^{-2}$ M$_\\odot]$. \\item at $M<5\\times10^{11}$ M$_\\odot$, a significant intrinsic evolution has been detected across the redshift range $z=2-3$, strongly dependent on mass. The dependence or the 5.8$\\mu$m-peakers number density on $\\left(1+z\\right)$ has powers of $\\sim-0.4$ and $\\sim-0.65$ for $M\\sim2\\times10^{11}$ and $4\\times10^{11}$ M$_\\odot$ respectively. In the highest mass bins ($M\\ge5\\times10^{11}$ M$_\\odot$) the number density of IR-peakers keeps nearly constant. \\item comparison to literature data for the $z=2-3$ mass function shows an overall agreement of the 5.8$\\mu$m-peaker comoving number density to that of the K20, MUNICS, and GOODS-MUSIC surveys, in the higher mass regime. At lower masses a significant evolutionary correction should be applied, when the mass function is averaged over a wide redshift range. \\item a significant evolution of the stellar mass function of $M\\gtrsim10^{11}$ M$_\\odot$ galaxies with respect to the local estimate was detected: $\\Phi(z=2-3)\\le0.1\\times\\Phi(z=0)$. Combining 5.8$\\mu$m-peakers and literature data \\citep[e.g.][]{fontana2006}, this implies that the bulk of massive galaxies was not yet in place by the time the Universe was $\\sim3$ Gyr old, but must have been assembled in the following $\\sim1.5$ Gyr of evolution. \\item current hydro-dynamical models significantly overestimate the number density of massive galaxies, while the semi-analytic approach underestimates it. \\item since SWIRE 5.8$\\mu$m-peakers sample only the very massive tail of the mass function, the Schechter slope $\\alpha$ cannot be constrained. Despite its very low significance, the best fit value $\\alpha=-0.30$ is similar to that found in the literature. The other best fit parameters are: $M^\\ast = 1.66 \\times10^{11}$ $[$h$_{70}^{-2}$ M$_\\odot]$ $\\pm$1 dex; $\\Phi^\\ast=0.00022^{+0.00004}_{-0.00009}$ $[$ h$_{70}^3$ Mpc$^{-3}]$. \\item the integrated stellar mass density of 5.8$\\mu$m peakers is $\\rho_\\star = 1.18 \\times 10^7$ $[$h$_{70}$ M$_\\odot$ Mpc$^{-3}]$, with a 3$\\sigma$ range of $\\pm 0.3$ dex. Only a lower limit could be set for 4.5$\\mu$m-peak galaxies: $\\rho_\\star\\ge6.55\\times10^6$ $[$h$_{70}$ M$_\\odot$ Mpc$^{-3}]$. \\item on average SWIRE massive 5.8$\\mu$m-peakers provide 30$-$50\\% of the total stellar mass density in galaxies at $z=2-3$. \\item 5.8$\\mu$m-peakers provide less than $\\sim10$\\% of the stellar mass locked in massive galaxies ($M=10^{11}-10^{12}$ M$_\\odot$) in the local Universe. \\end{itemize} The analysis carried out on SWIRE massive galaxies at $z>1.5$ over one square degree, highlighted the complementarity of wide-shallow and deep pencil-beam surveys. On one hand, sampling the faint end of the luminosity function, i.e. the low-mass end of the mass function, is needed in order to constrain the shape of the mass function and the total stellar mass density in galaxies at high redshift. On the other hand, very massive galaxies are rare objects on the sky, with a number density $1.2\\times10^{-4}$ $[$h$_{70}^3$ Mpc$^{-3}]$ and it is necessary to explore large volumes of Universe in order to fully characterize them. The natural extension of this analysis is to build the stellar mass function of high-$z$ galaxies over the whole SWIRE 49 deg$^2$ area, taking advantage of what we have learned thanks to the full multi-wavelength coverage in ELAIS-S1. At the time being, not much information is known about the environment hosting these massive high-$z$ objects. Further analyses of this population should be carried out probing also the environmental frame they belong to, in order to correctly interpret their role in the ``downsizing'' scenario." }, "0710/0710.2902_arXiv.txt": { "abstract": "We report on our initial analysis of a deep 510 ks observation of the Galactic oxygen-rich supernova remnant (SNR) G292.0+1.8 with the {\\it Chandra X-ray Observatory}. Our new {\\it Chandra} ACIS-I observation has a larger field of view and an order of magnitude deeper exposure than the previous {\\it Chandra} observation, which allows us to cover the entire SNR and to detect new metal-rich ejecta features. We find a highly non-uniform distribution of thermodynamic conditions of the X-ray emitting hot gas that correlates well with the optical [O {\\small III}] emission, suggesting the possibility that the originating supernova explosion of G292.0+1.8 was itself asymmetric. We also reveal spectacular substructures of a torus, a jet, and an extended central compact nebula all associated with the embedded pulsar J1124$-$5916. ", "introduction": "Supernova Remnant} Figure~\\ref{fig:fig1} shows an X-ray color image of G292.0+1.8 created from our deep {\\it Chandra} data. The outer boundary of the radio SNR is overlaid with a white contour. The energy bands displayed in each color were chosen to emphasize major atomic line emission that illustrates the distribution of electron temperatures and ionization states across the SNR (Table~\\ref{tbl:tab1}). We note that line and underlying continuum emission is included in each subband; e.g., the bright blue emission near the SNR's center is dominated by synchrotron emission from the PWN (see the inset in Figure~\\ref{fig:fig1}; Hughes et al.\\ 2001). Our deep ACIS-I exposure comprehensively images the entire SNR to its faint outermost edge, which matches well the extent of the radio remnant (Figure~\\ref{fig:fig1}) and traces the location of the blast wave. Ejecta knots (Park et al.\\ 2004, Gonzalez \\& Safi-Harb 2003) appear brightly colored -- yellow, green or blue -- in Figure~\\ref{fig:fig1}. They extend closest to the rim in the west and south and are furthest away in the SE quadrant. Due to their red-orange color in Figure~\\ref{fig:fig1} and positional coincidence with [O {\\small III}] ejecta \\citep{wink06}, the southernmost X-ray knots are likely to be O-rich as well. The SNR's full diameter is $\\sim$9$\\farcm$6 (N-S) and $\\sim$8$\\farcm$4 (E-W), corresponding to physical sizes of $\\sim$14.7$-$16.7 $d_6$ pc, where $d_6$ is the distance to G292.0+1.8 in units of 6 kpc, the distance we assume throughout \\citep{gaen03}. Our new image of G292.0+1.8 shows little evidence for the featureless, spectrally-hard, and geometrically-thin X-ray filaments that trace sites of efficient particle acceleration in other young SNRs, such as Cas A, Tycho, and SN 1006. Instead the SNR's rim is defined by spectrally-soft emission that is faint and diffuse in places (e.g., the SE rim) and filamentary elsewhere. Particle acceleration is evidently occurring under rather different conditions in G292.0+1.8 than other young Galactic SNRs. The SNR interior contains a complex network of knots and filaments with a variety of colors and morphologies. The overall color distribution of these features is highly asymmetric; red-orange emission is dominant in the S-SE, while the W-NW regions are bright in emission appearing green-blue in color. We are confident that these variations largely reflect differences in the underlying distributions of gas temperature and ionization in the metal-rich ejecta, rather than variable foreground absorption. This is supported by the relative spatial distributions of the Ne He$\\alpha$ and Ly$\\alpha$ lines. Because these lines are separated in energy by only $\\sim$0.1 keV, their flux ratio is insensitive to absorption variations. The equivalent width maps show that Ne He$\\alpha$ emission is relatively enhanced in the S-SE, while Ne Ly$\\alpha$ emission is enhanced in the W-NW \\citep{park02}. The metal abundance variation is also unlikely a significant contributor for the observed large-scale color variation based on our hardness ratio (HR) analysis (see below). We constructed a HR map (Figure~\\ref{fig:fig2}a) to investigate further the variation in the thermodynamic state of the ejecta across G292.0+1.8. (For technical reasons we added the continuum component [$E$ $<$ 1.25 keV] to the soft band line emission to enhance the hard band lines and suppress the strong hard continuum emission from the PWN.) There are prominent enhancements in Figure~\\ref{fig:fig2}a near the W and NW boundary of the SNR that generally trace ``fingers'' of hot ejecta protruding out to the very boundary of the SNR (green-blue filaments in Figure~\\ref{fig:fig1}). There is also an overall large-scale variation from HR values $\\sim$2.4 in the hard ``NW'' region to $\\sim$1.1 in the soft ``SE'' region (Figure~\\ref{fig:fig2}a). Example spectra extracted from small regions within the larger ``NW'' and ``SE'' regions clearly show the distinctive line ratios responsible for the observed HR variation (Figure~\\ref{fig:fig2}b). We can draw on our previous work fitting the spectra of individual knots to relate HR values to plasma temperatures. Spectral region 3 from Park et al. (2004) lies in the hard ``NW'' region, and spectral analysis yields a best-fit $kT \\sim 5$ keV, corresponding to HR $\\sim$ 2.5. We find that this HR value strongly traces the electron temperature rather than on individual elemental abundances. Although we did not study knots in the soft ``SE'' region previously, our preliminary spectral modeling of the SE regions favors significantly lower temperatures ($kT$ $\\sim$ 0.7 keV), corresponding to HR values $\\sim$ 1.0. Thus, the observed HR distribution reveals a large-scale spatial variation of thermal condition of metal-rich ejecta, which could not be detected by previous low-resolution data \\citep{hughes94}; i.e., a significantly higher temperature of the hot gas ($kT$ $\\sim$ 5 keV) is implied in the N-W regions, while a relatively lower-temperature plasma ($kT$ $\\la$ 1 keV) prevails in the S-E regions of the SNR. There is also a very similar and highly significant gradient in the optical properties of G292.0+1.8. The region with the lowest HR values is coincident with the bulk of the optical [O {\\small III}] emission indicated by ``SE'' region in Figure~\\ref{fig:fig2}a \\citep{ghav05,wink06}. Across the projected middle of the SNR (within an $\\sim$3$'$ wide region aligned roughly N-S), there are dozens of isolated optical knots (generally uncorrelated with X-ray knots), while on the western edge no optical emission appears at all. This high degree of correlation between the optical and X-ray properties suggests the possibility that radiative cooling in the ejecta is responsible for the SE-NW gradient in observed properties. In this picture, the SE ejecta would be undergoing significant, perhaps catastrophic cooling. Across the projected middle of the SNR, cooling is probably just beginning to occur in isolated knots that happen to have the appropriate thermodynamic conditions, while the emission toward the NW remains fully nonradiative. Variation in the ambient density surrounding G292.0+1.8 could provide an explanation for the observed asymmetry in the ejecta properties. However, there is no evidence for a higher density in the SE regions \\citep{braun86} as would be expected. In fact, 60 $\\mu$m images\\footnote{The apparent differences in the detailed structure of the two IR images in Figure~\\ref{fig:fig1} are unlikely to be real because of the highly processed nature of both images. The {\\it ISO} image was constructed from a rastor scan of many pointings using a $3\\times 3$ pixel ($\\sim$45$^{\\prime\\prime}$ per pixel) IR array that, unfortunately, did not fully cover the SE rim of the SNR. The {\\it IRAS} image, which did cover the entire SNR, was the result of an image reconstruction process intended to improve the angular resolution from the native value of several arcminutes to the level of 1$^{\\prime}$$-$2$^{\\prime}$. Given these significant differences, the overall general agreement between the images, specifically, the brightness of the SW rim and the faintness toward the SE, seems reasonable to us.} (see upper right and upper left insets to Figure~\\ref{fig:fig1}) show that around the rim of the SNR the SE region is a minimum in flux, while the SW is a maximum. Thus, albeit somewhat speculative at the current stage of the analysis, we raise the possibility that the ejecta asymmetry has its origin in some intrinsic asymmetry of the SN explosion itself, such as a variation in the density or velocity distribution from SE to NW. Further work including detailed X-ray spectral analysis of ejecta and hydrodynamic studies appropriate for G292.0+1.8 are required to test this asymmetric explosion scenario. ", "conclusions": "" }, "0710/0710.5525_arXiv.txt": { "abstract": "Temperature anisotropies in the Cosmic Microwave Background (CMB) are affected by the late Integrated Sachs-Wolfe (lISW) effect caused by any time-variation of the gravitational potential on linear scales. Dark energy is not the only source of lISW, since massive neutrinos induce a small decay of the potential on small scales during both matter and dark energy domination. In this work, we study the prospect of using the cross-correlation between CMB and galaxy density maps as a tool for constraining the neutrino mass. On the one hand massive neutrinos reduce the cross-correlation spectrum because free-streaming slows down structure formation; on the other hand, they enhance it through their change in the effective linear growth. We show that in the observable range of scales and redshifts, the first effect dominates, but the second one is not negligible. We carry out an error forecast analysis by fitting some mock data inspired by the Planck satellite, Dark Energy Survey (DES) and Large Synoptic Survey Telescope (LSST). The inclusion of the cross-correlation data from Planck and LSST increases the sensitivity to the neutrino mass $m_{\\nu}$ by 38\\% (and to the dark energy equation of state $w$ by 83\\%) with respect to Planck alone. The correlation between Planck and DES brings a far less significant improvement. This method is not potentially as good for detecting $m_{\\nu}$ as the measurement of galaxy, cluster or cosmic shear power spectra, but since it is independent and affected by different systematics, it remains potentially interesting if the total neutrino mass is of the order of 0.2~eV; if instead it is close to the lower bound from atmospheric oscillations, $m_{\\nu} \\sim 0.05$~eV, we do not expect the ISW-galaxy correlation to be ever sensitive to $m_{\\nu}$. ", "introduction": "As photons pass through a changing gravitational potential well, they experience a redshift or a blueshift, depending on whether the well grows or decays respectively. Cosmic microwave background (CMB) photons can experience such variations between the time of last scattering and their detection now. This effect was first described by Sachs and Wolfe in 1967~\\cite{Sachs:1967er}, and hence is dubbed the integrated Sachs-Wolfe effect (ISW). During a Cold Dark Matter (CDM) and/or baryon dominated era, the gravitational potential distribution remains frozen, and the ISW effect has no net effect on the blackbody temperature of CMB photons. This property is crucially related to the fact that non-relativistic matter (like CDM and baryons) has a vanishing sound speed, and experiences gravitational clustering on all sub-Hubble scales after photon decoupling, as described by the Poisson equation. In such a situation, the universal expansion and the gravitational contraction compensate each other in such a way as to maintain a static gravitational potential. However, when the expansion rate is affected by any type of matter with a non-vanishing sound speed, e.g. during Dark Energy (DE) domination, the gravitational perturbations decay and the cosmic photon fluid experiences a blue shift, acquiring extra temperature perturbations related to the intervening pattern of matter perturbations. It was first proposed by Crittenden and Turok in 1995~\\cite{Crittenden:1995ak} to cross correlate maps of temperature perturbations in the CMB with those of matter overdensities in large scale structures (LSS), in order to measure a possible acceleration of the universe's expansion. However, the CMB and LSS data available at that time were not good enough for such an ambitious goal, and the first strong indication of a positive acceleration came in 1998 from the side of type-Ia supernovae~\\cite{Perlmutter:1998np,Riess:1998cb}. Analyses of the first (2003) and second (2006) data releases of the Wilkinson Microwave Anisotropy Probe (WMAP)~\\cite{Spergel:2003cb,Spergel:2006hy} were the first to indicate the existence of Dark Energy independent of acceleration, by means of the location of the second peak in the CMB power spectrum. Simultaneously, a number of interesting papers presented the first detections of the ISW effect by cross-correlating WMAP anisotropy maps with various LSS data sets ~\\cite{Fosalba:2003ix,Boughn:2003yz,Fosalba:2003iy,Afshordi:2003xu, Padmanabhan:2004,Cabre:2006qm,Cabre:2006uj,Giannantonio:2006du,McEwen:2007}, now able to give an independent measure for the acceleration of the expansion of the universe. The domination of Dark Energy is not the only source of gravitational potential evolution and of a net ISW effect. On small cosmological scales, as soon as matter perturbations exceed the linear regime, gravitational perturbations start to grow and to redshift CMB photons. This effect, called the Rees-Sciama effect, has not been significantly detected until now~\\cite{Puchades:2006gs}. CMB photons can also be scattered by gravitational lensing \\cite{SeljakZaldarriaga:2000} and by the Sunyaev-Zeldovich (SZ) effect \\cite{SZ:1969} (see \\cite{Fosalba:2003iy,Afshordi:2003xu,Afshordi:2007} for detections in CMB-LSS cross-correlation analysis). An other party expected to affect the evolution of gravitational perturbations --at least by a small amount-- is the background of massive neutrinos. Over thirty years ago massive neutrinos were proposed as a Hot Dark Matter (HDM) candidate, and later ruled out as the dominant dark component, since HDM tends to wash out small scale overdensities during structure formation~\\cite{Primack:2001ib}. Observed neutrino oscillations however constrain neutrinos to have a mass~\\cite{Maltoni:2004ei,Fogli:2005cq}. In addition, the presence of a Cosmic Neutrino Background (CNB) is strongly suggested on the one hand by the abundance of light elements produced during primordial nucleosynthesis~\\cite{Cuoco:2003cu,Steigman:2005uz,Mangano:2006ur}, and on the other hand by CMB anisotropies~\\cite{Crotty:2003th,Pierpaoli:2003kw,Barger:2003zg,Trotta:2004ty,Hannestad:2005jj,Hannestad:2006mi,Ichikawa:2006vm,Ichikawa:2006vm,deBernardis:2007bu,Hamann:2007pi}. Therefore, a small fraction of HDM is expected to coexist with the dominant CDM component. On small cosmological scales (for instance, cluster scale), the free-streaming of massive neutrinos should induce a slow decay of gravitational and matter perturbations~\\cite{Bond:1980ha}, acting during both matter and Dark Energy domination. This effect depends on the total neutrino mass summed over all neutrino families, $m_{\\nu}=\\sum_i m_i$, unlike laboratory experiments based on tritium decay or neutrinoless double-beta decay, which probe different combinations: hence, a cosmological determination of the total neutrino mass would bring complementary information to the scheduled particle physics experiments~\\cite{Hannestad:2006zg,Lesgourgues:2006nd}. % The free streaming of massive neutrinos has not yet been detected~\\cite{Hannestad:2007tu}, but there are good prospects to do so in the future, since the smallest total neutrino mass allowed by data on atmospheric neutrino oscillations ($m_{\\nu} \\geq \\sqrt{\\Delta m^2_{\\rm atm}} \\sim 0.05$~eV) implies at least a 5\\% suppression in the matter/gravitational small-scale power spectrum~\\cite{Hannestad:2006zg,Lesgourgues:2006nd}. A positive detection --even in the case of minimal mass-- could follow from the analysis of future galaxy/cluster redshift surveys~\\cite{Lesgourgues:2004ps,Wang:2005vr,Hannestad:2007cp}, weak lensing surveys~\\cite{Song:2003gg,Hannestad:2006as}, Lyman-$\\alpha$ forest analysis, cluster counts~\\cite{Wang:2005vr}, etc. The goal of measuring the neutrino mass from cosmology is very ambitious since each of these methods suffers from its own source of systematics (bias issues, modeling of non-linear clustering, ...). Therefore, a robust detection could only be achieved by comparing the results from various types of experiments. The goal of this work is to describe a possible cosmological determination of the absolute neutrino mass scale through the ISW effect induced by neutrino free-streaming on CMB temperature maps, using as an observable the cross-correlation function of galaxy-temperature maps. This possibility was investigated previously by Ichikawa and Takahashi \\cite{Ichikawa:2005hi} (and suggested again recently in \\cite{Kiakotou:2007pz}). As neutrinos slow down the growth of structure, we expect the blueshift caused by an accelerated expansion to be more pronounced if neutrinos have a larger mass. On the other hand, the distribution of matter inducing the late ISW effect is smoother in case of free-streaming by massive neutrinos. These two antagonist effects should in principle induce some mass-dependent variations in the galaxy-temperature cross-correlation function. In section~\\ref{sec:theory} of this paper we give an outline of the theory of the ISW-effect in the presence of a neutrino mass. In section~\\ref{sec:mock}, we use some mock data with properties inspired from the Planck satellite, Dark Energy Survey (DES) and Large Synoptic Survey Telescope (LSST) in order to show the potential impact of this method in the future. ", "conclusions": "We have studied here the possibility to use the cross-correlation between CMB and galaxy density maps as a tool for constraining the neutrino mass. On one hand massive neutrinos reduce the cross-correlation spectrum because their free-streaming slows down structure formation; on the other hand, they enhance it because of the behavior of the linear growth in presence of massive neutrinos. Using both analytic approximations and numerical computations, we showed that in the observable range of scales and redshifts, the first effect dominates, but the second one is not negligible. Hence the cross-correlation between CMB and LSS maps could bring some independent information on neutrino masses. We performed an error forecast analysis by fitting some mock data inspired from the Planck satellite, Dark Energy Survey (DES) and Large Synoptic Survey Telescope (LSST). For Planck and LSST, the inclusion of the cross-correlation data increases the sensitivity to $m_{\\nu}$ by 38\\%, $w$ by 83\\% and $\\Omega_{dm} h^2$ by 30\\% with respect to the CMB data alone. With the fiducial model employed in this analysis (based on eight free parameters) the standard deviation for the neutrino mass is equal to 0.38~eV for Planck alone and 0.27~eV for Planck plus cross-correlation data. This is far from being as spectacular as the sensitivity expected from the measurement of the auto-correlation power spectrum of future galaxy/cluster redshift surveys or cosmic shear experiments, for which the predicted standard deviation is closer to the level of 0.02~eV, leading to a 2$\\sigma$ detection even in the case of the minimal mass scenario allowed by current data on neutrino oscillations (see \\cite{Lesgourgues:2006nd} for a review). However, the method proposed here is independent and affected by different systematics. So, it remains potentially interesting, but only if the neutrino mass is not much smaller than $m_{\\nu} \\sim 0.2$~eV." }, "0710/0710.0907_arXiv.txt": { "abstract": "This paper presents the results of a {\\em Spitzer} IRAC $3-8$ $\\mu$m photometric search for warm dust orbiting 17 nearby, metal-rich white dwarfs, 15 of which apparently have hydrogen dominated atmospheres (type DAZ). G166-58, G29-38, and GD 362 manifest excess emission in their IRAC fluxes and the latter two are known to harbor dust grains warm enough to radiate detectable emission at near-infrared wavelengths as short as 2 $\\mu$m. Their IRAC fluxes display differences compatible with a relatively larger amount of cooler dust at GD 362. G166-58 is presently unique in that it appears to exhibit excess flux only at wavelengths longer than about 5 $\\mu$m. Evidence is presented that this mid-infrared emission is most likely associated with the white dwarf, indicating that G166-58 bears circumstellar dust no warmer than $T\\sim400$ K. The remaining 14 targets reveal no reliable mid-infrared excess, indicating the majority of DAZ stars do not have warm debris disks sufficiently opaque to be detected by IRAC. ", "introduction": "The {\\em Spitzer Space Telescope} opened a new phase space to white dwarf researchers interested in the infrared properties of degenerate stars and their environments. The majority of white dwarfs are inaccessible from the ground beyond 2.4 $\\mu$m due to their intrinsic faintness combined with the ever-increasing sky brightness towards longer wavelengths \\citep*{gla99}. This limits any white dwarf science which aims to study matter radiating at $T<1500$ K. Prior to the launch of {\\em Spitzer}, only one previously published, directed mid-infrared study of white dwarfs exists ; an {\\em Infrared Space Observatory} search for dust emission around 11 nearby white dwarfs, 6 of which have metal-rich photospheres \\citep*{cha99}. Owing to the superb sensitivity of {\\em Spitzer} \\citep*{wer04}, a Cycle 1 IRAC program was undertaken to search for warm dust emission associated with cool, hydrogen atmosphere white dwarfs with photospheric metals, the DAZ stars. This paper presents a synopsis of the IRAC results, including the detection of $5-8$ $\\mu$m flux excess at G166-58, $3-8$ $\\mu$m data on G29-38 and GD 362, and also includes Gemini $3-4$ $\\mu$m spectroscopy of G29-38. ", "conclusions": "Although a survey of only 17 stars does not allow robust statistics, it is clear that the majority of DAZ white dwarfs do not harbor warm dust of sufficient emitting surface area to be detected with IRAC. If most or all of these stars do host circumstellar material, the fractional luminosities must be relatively low compared to currently known dusty white dwarfs. This could result from a modest amount of dust (as in the zodiacal cloud), large particle sizes, or cooler material further from the star. Another possibility is that any warm dust produced within the Roche limit of a white dwarf is swiftly destroyed through mutual collisions, not unlike ice in a blender, as it orbits with Keplerian velocities near $0.003c$. In optically thin disks, particles with orbital period $p$ will collide on a timescale given by $t_{\\rm coll}=p/\\tau$. The tidal rings at G29-38 and GD 362 have been modeled to extend from approximately $0.2-0.4$ $R_{\\odot}$, where a typical orbital period is only $p=0.6$ hr and the resulting collision timescale for $\\tau\\sim0.01$ is $t_{\\rm coll}=2.5$ dy. If a sizeable fraction of dust produced in a tidal disruption event is initially optically thin, then both collisions and Poynting-Robertson drag will compete to quickly annihilate this material. The ratio of these two timescales for dust particles orbiting a distance $D$ from a star of mass $M$ can be written as \\begin{equation} \\gamma = \\frac{t_{\\rm pr}}{t_{\\rm coll}} = 693 \\ \\frac{\\sqrt{MD} \\rho a\\tau}{LQ_{\\rm pr}} \\end{equation} \\noindent where $M$ and $L$ are in solar units, $D$ is in AU, $\\rho$ is in g ${\\rm cm}^{-3}$ , and $a$ is in microns. This fraction reaches a minimum for 0.1 $\\mu$m grains with 1 g ${\\rm cm}^{-3}$ at the inner disk edge, and yields $\\gamma_{\\rm min}>10$ for all possible white dwarf disk parameters. Table \\ref{tbl6} gives minimum values for $\\gamma$ at the inner edges of the disks at G29-38, GD 362, and G166-58. Therefore collisions will erode dust grains faster than they can be removed by angular momentum loss. This is also true in the case of optically thick disks where: 1) the bulk of material is shielded from starlight, and hence the Poynting-Robertson effect is diminished, and 2) the collision timescale is less than half the orbital period \\citep*{esp93}. Hence, for a wide range of disk densities, it is plausible that mutual collisions within an evolving dust ring at a typical white dwarf will result in the relatively rapid self-annihilation of the micron size grains required to radiate efficiently at $3-30$ $\\mu$m. Following the tidal disruption of an asteroid, if one models the dust produced as a collisional cascade, the expected particle size distribution behaves classically as $n(a)\\propto a^{-3.5}$ \\citep*{doh69}. For dust at main sequence stars, this distribution is reshaped on short timescales as sub-micron size grains are removed by radiation pressure which, as shown above, does not apply in the case of white dwarfs. In the absence of radiation pressure, the average particle in will have a size $\\bar{a}= 5/3$ $a_{\\rm min}$. For practical purposes, at white dwarfs one can assume $a_{\\rm min}=0.01$ $\\mu$m, where particles are already inefficient absorbers and emitters of infrared radiation, and anything smaller approaches the size of gas molecules and atoms. With such a distribution of extant dust, 99.7\\% of the particles will have sizes $a\\leq0.1$ $\\mu$m, leaving a paltry fraction of larger particles which could effectively support infrared emission from the disk. Clearly, such small particles and gas might be present at most or all white dwarfs which show signs of circumstellar accretion such as the DAZ stars. On the other hand, the persistence of warm dust disks at several white dwarfs \\citep*{jur07b} must be explained despite the fact that in some or most cases where it is produced, it may also be efficiently destroyed. One possibility is that the disk density (which could contain gas) becomes sufficiently high as to damp out collisions in the disk and also make it optically thick, thus somewhat protecting it from self-erosion and drag forces simultaneously. The evolution of such a dense, fluid-like ring is then dominated by viscous forces (differential rotation and random motions) which cause it to spread, losing energy in the process \\citep*{esp93}. The maximum lifetime of such a ring occurs at minimum viscosity \\begin{equation} t_{\\rm ring} \\approx \\frac{ {{\\rho}^2} {w^2} p} {2 \\pi {{\\sigma^2}} } \\end{equation} \\noindent where $w$ is the radial extent of the ring and $\\sigma$ is the surface mass density \\citep*{esp93}. If the mass of a large solar system asteroid, $10^{24}$ g, were spread into a tidal ring of negligible height ($h<10$ m), a radial extent $0.2-0.4$ $R_{\\odot}$, consisting of micron size particles orbiting a typical white dwarf, the resulting volume mass density (0.55 g ${\\rm cm}^{-3}$) would be sufficiently high that the mean free path of particles is on the same order as their size. This could effectively damp out collisions, thus minimizing viscosity, and with a resulting surface mass density of $\\sigma\\approx550$ g ${\\rm cm}^{-2}$, permit a potential disk lifetime -- in the absence of competing forces -- longer than the Gyr white dwarf cooling timescales. However, for sustained accretion rates as low as $\\dot{M}=10^{10}$ g ${\\rm s}^{-1}$, a $10^{24}$ g disk would become fully consumed within several Myr. Large rocks and colder material orbiting at $D\\ga100$ $R_{\\odot}$ will be unaffected by any of the aforementioned processes, and such a reservoir of material is strictly necessary to supply some fraction of DAZ white dwarfs with photospheric metals, regardless of circumstellar dust production (collisional versus tidal) and evolution (persistence versus destruction). The overall number of white dwarfs with remnant planetesimal belts may be rather high based on a growing number of detections. If one takes 12\\% (\\S4.1) as the fraction of DAZ stars with circumstellar dust as observed by {\\em Spitzer} to date, 20\\% as the fraction of DAZ stars among cool DA white dwarfs \\citep*{zuc03}, and 80\\% as the number of cool DA stars among all white dwarfs in the field \\citep*{eis06}, then a lower limit to the number of white dwarfs with asteroid-type belts is at least 2\\%. This fraction could be as high as 20\\% if the majority of metal-rich white dwarfs harbor circumstellar matter, which raises important questions about the implied frequency of planetesimal belts around main-sequence stars and the current detection rate (see the Appendix of \\citealt*{jur06}). Owing to their low luminosities, white dwarfs which may have been polluted by heavy elements in winds or transferred material from substellar companions \\citep*{deb06,dob05,zuc03,sio84} are easily identified with IRAC observations \\citep*{mul07,han06,far05a,far05b} down to T dwarf temperatures. There is no evidence of such companions in the data presented here, ruling out all but the coldest brown dwarfs, active planets and moons as close orbiting, companion-like polluters (Farihi et al. 2008, in preparation)." }, "0710/0710.5180_arXiv.txt": { "abstract": "Mass bounds on dark matter (DM) candidates are obtained for particles that decouple in or out of equilibrium while ultrarelativistic with {\\bf arbitrary} isotropic and homogeneous distribution functions. A coarse grained Liouville invariant primordial phase space density $ \\mathcal D $ is introduced which depends solely on the distribution function at decoupling. The density $ \\mathcal D $ is explicitly computed and combined with recent photometric and kinematic data on dwarf spheroidal satellite galaxies in the Milky Way (dShps) and the observed DM density today yielding upper and lower bounds on the mass, primordial phase space densities and velocity dispersion of the DM candidates. Combining these constraints with recent results from $N$-body simulations yield estimates for the mass of the DM particles in the range of a few keV. We establish in this way a direct connection between the microphysics of decoupling \\emph{in or out} of equilibrium and the constraints that the particles must fulfill to be suitable DM candidates. If chemical freeze out occurs before thermal decoupling, light bosonic particles can Bose-condense. We study such Bose-Einstein {\\it condensate} (BEC) as a dark matter candidate. It is shown that depending on the relation between the critical ($ T_c $) and decoupling ($ T_d $) temperatures, a BEC light relic could act as CDM but the decoupling scale must be {\\it higher} than the electroweak scale. The condensate hastens the onset of the non-relativistic regime and tightens the upper bound on the particle's mass. A non-equilibrium scenario which describes particle production and partial thermalization, sterile neutrinos produced out of equilibrium and other DM models is analyzed in detail and the respective bounds on mass, primordial phase space density and velocity dispersion are obtained. Thermal relics with $ m \\sim \\mathrm{few}\\,\\mathrm{keV} $ that decouple when ultrarelativistic and sterile neutrinos produced resonantly or non-resonantly lead to a primordial phase space density compatible with {\\bf cored} dShps and disfavor cusped satellites. Light Bose-condensed DM candidates yield phase space densities consistent with {\\bf cores} and if $ T_c\\gg T_d $ also with cusps. Phase space density bounds on particles that decoupled non-relativistically combined with recent results from N-body simulations suggest a potential tension for WIMPs with $ m \\sim 100\\,\\mathrm{GeV},T_d \\sim \\,10\\,\\mathrm{MeV} $. ", "introduction": "Although the existence of dark matter (DM) was inferred several decades ago \\cite{zwoo}, its nature still remains elusive. Candidate dark matter particles are broadly characterized as cold, hot or warm depending on their velocity dispersions. The clustering properties of collisionless DM candidates in the linear regime depend on the free streaming length, which roughly corresponds to the Jeans length with the particle's velocity dispersion replacing the speed of sound in the gas. Cold DM (CDM) candidates feature a small free streaming length favoring a bottom-up hierarchical approach to structure formation, smaller structures form first and mergers lead to clustering on the larger scales. Among the CDM candidates are weakly interacting massive particles (WIMPs) with $m \\sim 10-10^{2}\\,\\mathrm{GeV}$. Hot DM (HDM) candidates feature large free streaming lengths and favor top down structure formation, where larger structures form first and fragment. HDM particle candidates are deemed to have masses in the few $\\mathrm{eV}$ range, and warm DM (WDM) candidates are intermediate with a typical mass range $ m \\sim 1-10 \\,\\mathrm{keV} $. \\medskip The \\emph{concordance} $ \\Lambda\\mathrm{CDM} $ standard cosmological model emerging from CMB, large scale structure observations and simulations favors the hypothesis that DM is composed of primordial particles which are cold and collisionless \\cite{primack}. However, recent observations hint at possible discrepancies with the predictions of the $ \\Lambda\\textrm{CDM} $ concordance model: the satellite and cuspy halo problems. \\\\ The satellite problem, stems from the fact that CDM favors the presence of substructure: much of the CDM is not smoothly distributed but is concentrated in small lumps, in particular in dwarf galaxies for which there is scant observational evidence so far. A low number of satellites have been observed in Milky-Way sized galaxies \\cite{kauff,moore,moore2,klyp}. This substructure is a consequence of the CDM power spectrum which favors small scales becoming non-linear first, collapsing in the bottom-up hierarchical manner and surviving the mergers as dense clumps \\cite{moore,klyp}. \\\\ The cuspy halo problem arises from the result of large scale $N$-body simulations of CDM clustering which predict a monotonic increase of the density towards the center of the halos \\cite{dubi,frenk,moore2,bullock,cusps}, for example the universal Navarro-Frenk-White (NFW) profile $ \\rho(r)\\sim r^{-1}(r+r_0)^{-2} $ \\cite{frenk} which describes accurately clusters of galaxies, but indicates a divergent cusp at the center of the halo. Recent observations seem to indicate central cores in dwarf galaxies \\cite{dalcanton1,van,swat,gilmore}, leading to the 'cusps vs cores' controversy. \\\\ A recent compilation of observations of dwarf spheroidal galaxies dSphs \\cite{gilmore}, which are considered to be prime candidates for DM subtructure \\cite{spergel}, seem to \\emph{favor} a core with a smoother central density and a low mean mass density $ \\sim 0.1\\,M_{\\odot}/\\mathrm{pc}^3 $ rather than a cusp \\cite{gilmore}. The data cannot yet rule out cuspy density profiles which allow a maximum density $ \\lesssim 60 \\,M_{\\odot}/\\mathrm{pc}^3 $ and the interpretation and analysis of the observations is not yet conclusive \\cite{dalcanton1,van2}. These \\emph{possible} discrepancies have rekindled an interest in WDM particles, which feature a velocity dispersion larger than CDM particles, and consequently larger free-streaming lengths which smooth-out the inner cores and would be prime candidates to relieve the cuspy halo and satellite problems \\cite{turok}. \\medskip A possible WDM candidate is a sterile neutrino \\cite{dw,este,kusenko} with a mass in the $\\mathrm{keV}$ range and produced via their mixing and oscillation with an active neutrino species either non-resonantly \\cite{dw}, or through MSW (Mikheiev-Smirnov-Wolfenstein) resonances in the medium \\cite{este}. Sterile neutrinos can decay into a photon and an active neutrino (more precisely the largest mass eigenstate decays into the lowest one and a photon) \\cite{pal} yielding the possibility of direct constraints on the mass and mixing angle from the diffuse X-ray background \\cite{Xray}. \\medskip Observations of cosmological structure formation via the Lyman-$\\alpha$ forest provide a complementary probe of primordial density fluctuations on small scales which yield an indirect constraint on the masses of WDM candidates. While constraints from the diffuse X-ray background yield an \\emph{upper} bound on the mass of a putative sterile neutrino in the range $ 3-8\\,\\mathrm{keV} $ \\cite{Xray}, the latest Lyman-$\\alpha$ analysis \\cite{lyman} yields \\emph{lower} bounds in the range $ 10-13\\,\\mathrm{keV} $ in tension with the X-ray constraints. More recent constraints from Lyman-$\\alpha$ yield a lower limit for the mass of a WDM candidate $ m_{WDM} \\gtrsim 1.2\\,\\mathrm{keV} \\,(2\\sigma) $ for an early decoupled \\emph{thermal} relic and $m_{WDM} \\gtrsim 5.6\\,\\mathrm{keV} \\,(2\\sigma)$ for sterile neutrinos \\cite{viel}. Strong upper limits on the mass and mixing angles of sterile neutrinos have been recently discussed \\cite{beacom}, however, there are uncertainties as to whether WDM candidates can explain large cores in dSphs \\cite{strigari}. It has been recently argued \\cite{palazzo} that if sterile neutrinos are produced non-resonantly \\cite{dw} the combined X-ray and Lyman-$\\alpha$ data suggest that these \\emph{cannot} be the \\emph{only} WDM component, with an upper limit for their fractional relic abundance $ \\lesssim 0.7 $. Recent \\cite{boyarski2} constraints on a radiatively decaying DM particle from the EPIC spectra of (M31) by XMM-Newton confirms this result and places a stronger lower mass limit $ m < 4\\,\\mathrm{keV} $. All these results suggest that DM could be a mixture of several components with sterile neutrinos as viable candidates. \\medskip {\\bf Motivation and goals:} Although the $\\Lambda\\mathrm{CDM}$ paradigm describes large scale structure formation remarkably well, the \\emph{possible} small scale discrepancies mentioned above motivate us to study new constraints that different dark matter components must fulfill to be suitable candidates. Cosmological bounds on dark matter components primarily focused on standard model neutrinos \\cite{bond,TG}, heavy relics that decoupled in local thermodynamic equilibrium (LTE) when non-relativistic \\cite{LW,kt,dominik} or \\emph{thermal ultrarelativistic relics} \\cite{madsen,madsenbec,madsenQ,salu,hogan}. More recently, cosmological precision data were used to constrain the (HDM) abundance of low mass particles \\cite{pastor, steen,raffelt,mena} assuming these to be thermal relics. \\medskip The main results of this article are: \\medskip (\\textbf{a:}) We consider particles that decouple {\\bf in or out of LTE} during the radiation dominated era with an \\emph{arbitrary} (but homogeneous and isotropic) distribution function. Particles which decouple being ultrarelativistic eventually become non-relativistic because of redshift of physical momentum. We establish a direct connection between the microphysics of decoupling \\emph{in or out of LTE} and the constraints that the particles must fulfill to be suitable DM candidates \\emph{in terms of the distribution functions at decoupling}. \\medskip (\\textbf{b:}) We introduce a primordial coarse grained phase space density $$ \\mathcal{D} \\equiv \\frac{n(t)}{\\big\\langle \\vec{P}^2_f \\big\\rangle^\\frac32} \\; , $$ where $ n(t) $ is the number of particles per unit physical volume and $\\Big\\langle\\vec{P}^2_f\\Big\\rangle$ is the average of the physical momentum with the distribution function of the decoupled particle. $ \\mathcal D $ is a Liouville invariant after decoupling and only depends on the distribution functions at decoupling. In the non-relativistic regime $ \\mathcal D $ is simply related to the phase densities considered in refs. \\cite{dalcanton1,TG,hogan,madsenQ} and can only {\\bf decrease} by collisionless phase mixing or self-gravity dynamics \\cite{theo}. In the non-relativistic regime we obtain \\be \\label{Dint} \\mathcal{D} = \\frac1{3^\\frac32 \\; m^4} \\; \\frac{\\rho_{DM}}{\\sigma^3_{DM}} \\ee where $ \\sigma_{DM} $ is the primordial one-dimensional velocity dispersion and $ \\rho_{DM} $ the dark matter density. Combining the result for the primordial phase space density $ \\mathcal D $ determined by the mass and the distribution function of the decoupled particles, with the recent compilation of photometric and kinematic data on dSphs satellites in the Milky-Way \\cite{gilmore} yields {\\bf lower} bounds on the DM particle mass $ m $ whereas {\\bf upper} bounds on the DM mass are obtained using the value of the observed dark matter density today. Therefore the combined analysis of observational data from (dSphs), N-body simulations and the present DM density allows us to establish both \\emph{upper and lower} bounds on the mass of the DM candidates. We thus provide a link between the microphysics of decoupling, the observational aspects of dark matter halos and the DM mass value. \\medskip (\\textbf{c:}) Recent $N$-body simulations \\cite{numQ} indicate that the phase-space density decreases a factor $ \\sim 10^2 $ during gravitational clustering. This result combined with eq.(\\ref{Dint}) and the observed values on dSphs satellites \\cite{gilmore} yield $$ m_{cored}\\sim \\frac2{g^\\frac14} \\; \\mathrm{keV} \\quad , \\quad m_{cusp} \\sim\\frac8{g^\\frac14} \\; \\mathrm{keV} \\; . $$ for the masses of \\emph{thermal relics} DM candidates, where `cored' and 'cusp' refer to the type of profile used in the dShps description and $ 1\\leq g \\leq 4 $ is the number of internal degrees of freedom of the DM particle. Wimps with masses $ \\sim 100 \\, \\textrm{GeV} $ decoupling in LTE at temperatures $ T_d \\sim 10\\,\\mathrm{MeV} $ lead to primordial phase space densities many orders of magnitude larger than those observed in (dSphs). The results of $N$-body simulations, which yield relaxation by $2-3$ orders of magnitude\\cite{numQ} suggest a potential tension for WIMPs as DM candidates. However, the $N$-body simulations in ref.\\cite{numQ} begin with initial conditions with values of the phase space density much lower than the primordial one. Hence it becomes an important question whether the enormous relaxation required from the primordial values to those of observed in dSphs can be inferred from numerical studies with suitable (much larger) initial values of the phase space density. \\medskip (\\textbf{d:}) We study the possibility that the DM particle is a light Boson that undergoes Bose-Einstein Condensation (BEC) prior to decoupling while still ultrarelativistic. (This possibility was addressed in \\cite{madsenbec}). We analyze in detail the constraints on such BEC DM candidate from velocity dispersion and phase space arguments, and contrast the BEC DM properties to those of the hot or warm thermal relics. \\medskip (\\textbf{e:}) Non-equilibrium scenarios that describe various possible WDM candidates are studied in detail. These scenarios describe particle production \\cite{boydata} and incomplete thermalization \\cite{dvd}, resonant \\cite{dw} and non-resonant \\cite{este} production of sterile neutrinos and a model recently proposed \\cite{strigari} to describe cores in dSphs. Our analysis of the DM candidates is based on their masses, statistics and properties at decoupling (being it in LTE or not). We combine observations on dSphs \\cite{gilmore} and $N$-body simulations \\cite{numQ}, with theoretical analysis using the non-increasing property of the phase space density \\cite{TG,dalcanton1,hogan,theo}. The results from the combined analysis of the primordial phase space densities, the observational data on dSphs \\cite{gilmore} and the $N$-body simulations in ref.\\cite{numQ} are the following: \\begin{itemize} \\item{\\textbf{(i):} conventional thermal relics, and sterile neutrinos produced resonantly or non-resonantly with mass in the range $ m \\sim \\mathrm{few}\\,\\mathrm{keV} $ that decouple when ultrarelativistic lead to a primordial phase space density of the same order of magnitude as in cored dShps and disfavor cusped satellites for which the data \\cite{gilmore} yields a much larger phase space density. } \\item{\\textbf{(ii):} CDM from wimps that decouple when non-relativistic with $ m \\gtrsim 100~\\mathrm{GeV} $ and kinetic decoupling at $ T_d \\sim 10~\\mathrm{MeV} $ \\cite{dominik} yield phase space densities at least eighteen to fifteen orders of magnitude [see eqs.(\\ref{rss}), (\\ref{cusp}) and (\\ref{denw})] larger than the typical average in dSphs \\cite{gilmore}. Results from $N$-body simulations, albeit with initial conditions with much smaller values of the phase space density, yield a dynamical relaxation by a factor $ 10^2-10^3 $ \\cite{numQ}. If these results are confirmed by simulations with larger initial values there may be a potential tension between the primordial phase space density for \\emph{thermal relics} in the form of WIMPs with $ m\\sim 100\\,\\mathrm{GeV}, \\; T_d \\sim 10 $ MeV and those observed in dShps.} \\item{\\textbf{(iii):} Light bosonic particles decoupled while ultrarelativistic and which form a BEC lead to phase space densities consistent with cores and also consistent with cusps if $ T_c/T_d \\gtrsim 10 $. However if these thermal relics satisfy the observational bounds, they must decouple when $ g_d \\; g^{-\\frac34} \\; (T_d/T_c)^\\frac98 > 130 $, namely {\\it above} the electroweak scale. } \\end{itemize} \\medskip Section II analyzes the { generic} dynamics of decoupled particles for { any} distribution function, { with or without} LTE at decoupling, and for {\\bf different} species of particles. In section III we consider light thermal relics which decoupled in LTE as DM components: fermions and bosons, including the possibility of a Bose-Einstein condensate. Section IV deals with coarse grained phase space densities which are Liouville invariant and the new bounds obtained with them by using the observational dSphs data and recent results from $N$-body simulations, bounds from velocity dispersion, and the generalized Gunn-Tremaine bound. In Section V we study the case of particles that decoupled { out of equilibrium} and the consequences on the dark matter constraints. Section VI summarizes our conclusions. ", "conclusions": "We have obtained new constraints on light DM candidates that decoupled while ultrarelativistic in or out of LTE in terms of their distribution functions. The only assumption is that these distribution functions are homogeneous and isotropic. A Liouville invariant coarse grained primordial phase space density is introduced that allows to combine phase space density arguments with a recent compilation of photometric and kinematic data on dSphs galaxies to yield \\emph{new constraints} on the mass, velocity dispersion and phase space density of DM candidates. The new constraint on the mass range is \\be \\frac{62.36~\\mathrm{eV}}{\\mathcal{D}^{\\frac{1}{4}}} \\; \\Bigg[10^{-4} \\; \\frac{\\rho}{\\sigma^3} \\; \\frac{\\big(\\mathrm{km}/\\mathrm{s}\\big)^3}{M_\\odot/\\mathrm{kpc}^3} \\Bigg]^{\\frac14} \\leq m \\leq 2.695 \\; \\frac{2 \\; g_d \\; \\zeta(3)}{g \\; \\int^\\infty_0 y^2 \\; f_{d }(y) \\; dy} \\; \\mathrm{eV} \\; , \\ee where the primordial phase space density is given by \\be \\mathcal{D} = \\frac{g}{2 \\; \\pi^2} \\frac{\\Bigg[\\int_0^\\infty y^2 \\; f_d(y) \\; dy \\Bigg]^{\\frac52}}{\\Bigg[\\int_0^\\infty y^4 \\; f_d(y) \\; dy \\Bigg]^{\\frac32}} \\; , \\ee $ f_d(p_c/T_d) $ is the distribution function at decoupling, $ g $ the number of internal degrees of freedom of the particle, and $ \\rho/\\sigma^3 $ is the phase space density obtained from observations. The \\emph{upper bound} arises from requesting that the DM candidate has a density $ \\leq \\rho_{DM} $ today, and the \\emph{lower} bound arises from requesting that the phase space density in halos $ \\rho/\\sigma^3 $ be \\emph{smaller} than or equal to the primordial phase space density of the collisionless non-relativistic (today) DM component $$ \\rho_{DM}/\\sigma^3_{DM}=3^\\frac32 \\; m^4 \\; \\mathcal{D} . $$ We have studied the consequences of Bose-Einstein condensation of light ultrarelativistic particles when chemical freeze out occurs well before kinetic decoupling at $ T_d < T_c $ with $ T_c $ the critical temperature below which a non-vanishing condensate fraction exists. We find that the presence of the condensate hastens the onset of the non-relativistic regime and that Bose-Einstein condensed particles can effectively act as a CDM component {\\it even} when they decoupled being ultrarelativistic. The reason for this unusual behavior is that the particles in the condensate all have vanishing velocity dispersion. For \\emph{thermal relics} we find \\be \\mathcal{D} = g \\times \\left\\{ \\begin{array}{l} 1.963\\times 10^{-3}~~\\mathrm{Fermions},~\\mu_d=0 \\\\ 3.657\\times 10^{-3}~~\\mathrm{Bosons~no~BEC} \\\\ 3.657\\times 10^{-3}\\,\\Big(\\frac{T_c}{T_d} \\Big)^{ \\frac{15}{2}}~~\\mathrm{Bosons~with~BEC},~T_c>T_d \\\\ 8.442\\times 10^{-2}\\, g_d\\,Y_\\infty~~\\mathrm{non-relativistic~Maxwell-Boltzmann.} \\end{array} \\right.\\label{LTDD2} \\ee The combination of data in ref. \\cite{gilmore} from dSphs when applied to \\emph{light thermal relics} yields the mass range \\bea && \\frac{444~ \\mathrm{eV} }{g^{\\frac14}} \\leq m \\leq \\frac{g_d}{g}~ 4.253 ~ \\mathrm{eV} ~~\\mathrm{fermions~with}~ \\mu_d=0 \\; ,\\nonumber \\\\ && \\frac{ 380~\\mathrm{eV}}{g^{\\frac14}} \\leq m \\leq \\frac{g_d}{g}~ 2.695 ~\\mathrm{eV} ~~\\mathrm{bosons~with} ~ \\mu_d=0 \\; \\mathrm{and~no~BEC} \\; \\, \\nonumber \\\\ && \\frac{380~\\mathrm{eV}}{g^{\\frac14}}\\Bigg[\\frac{T_d}{T_c}\\Bigg]^\\frac{15}{8} \\leq m \\leq \\frac{g_d}{g}~ 2.695 ~\\Bigg[\\frac{T_d}{T_c}\\Bigg]^3 \\; \\mathrm{eV} ~~\\mathrm{BEC} \\; . \\label{massrangeF2} \\eea with the implication that if these particles are suitable DM candidates, they must decouple at high temperature when the effective number of ultrarelativistic degrees of freedom is $ g_d > 100 $. Namely, in absence of a BEC, thermal decoupling must occur above the electroweak scale. In the BEC case, for $ T_d\\ll T_c $, the fulfillment of the bound requires very large $ g_d $. Namely, in the presence of a BEC thermal decoupling occurs at a scale much larger than the electroweak scale for $ T_d\\ll T_c $. Assuming that the DM particle is the only component with the density $ \\rho_{DM} $ today, we obtained an independent bound from velocity dispersion which for the favored cored profiles \\cite{gilmore} yield the lower mass bound \\be \\frac{m}{\\mathrm{keV}} \\geq \\frac{0.855}{g^{\\frac13}_d}\\, \\Bigg[ \\frac{\\int_0^\\infty y^4 \\; f_d(y) \\; dy}{\\int_0^\\infty y^2 \\; f_d(y) \\; dy}\\Bigg]^{\\frac12} \\label{masdisp2} \\; . \\ee For light thermal relics this bound implies that $ m \\gtrsim 0.6-1.5\\,\\mathrm{keV} $ with a suppression factor $ T_d/T_c $ in the BEC case. For light thermal relics that decoupled while ultrarelativistic we find the primordial phase space density \\be \\frac{\\rho_{DM}}{\\sigma^3_{DM}} \\sim 10^6 ~\\frac{ \\mathrm{eV}/\\mathrm{cm}^3} {\\Big( \\mathrm{km}/\\mathrm{s} \\Big)^3}~\\Bigg( \\frac{m}{\\mathrm{keV}}\\Bigg)^3 \\; g_d \\; \\Bigg\\{\\begin{array}{l} 0.177~~~\\mathrm{Fermions} \\\\ 0.247~~~\\mathrm{Bosons ~ without ~ BEC} \\\\ 0.247\\,(T_c/T_d)^\\frac92 ~~~\\mathrm{Bosons ~ with ~ BEC} \\; . \\end{array} \\ee An enhancement factor $ (T_c/T_d)^\\frac92 $ appears in the r.h.s. in the presence of a BEC. For wimps with kinetic decoupling temperature $ 10 $MeV \\cite{dominik}, we find \\be\\label{wimpo} \\frac{\\rho_{wimp}}{\\sigma^3_{wimp}} \\sim 10^{24} \\; \\frac{ \\mathrm{eV}/\\mathrm{cm}^3}{\\Big( \\mathrm{km}/\\mathrm{s} \\Big)^3}~\\Bigg( \\frac{m}{100\\,\\mathrm{GeV}}\\Bigg)^3 \\; g_d \\; . \\ee The observational data compiled in ref. \\cite{gilmore} assuming a favored cored profile suggests \\be \\left(\\frac{\\rho_s}{\\sigma^3_s}\\right)_{cored} \\sim 5\\times 10^6 ~\\frac{ \\mathrm{eV}/\\mathrm{cm}^3} {\\Big( \\mathrm{km}/\\mathrm{s}\\Big)^3}\\,. \\label{core2} \\ee If the distribution of dark matter is cusped, ref. \\cite{gilmore} gives the value for the density $ \\rho_s \\sim 2 \\; \\mathrm{TeV}/\\mathrm{cm}^3 $ yielding \\be \\left(\\frac{\\rho_s}{\\sigma^3_s}\\right)_{cusped} \\sim 2\\times 10^9 ~\\frac{ \\mathrm{eV}/\\mathrm{cm}^3} {\\Big( \\mathrm{km}/\\mathrm{s} \\Big)^3} \\,.\\label{cusp2} \\ee Therefore, for $ g_d \\gtrsim 10 $ the primordial phase space density for \\emph{thermal relics} with $ m \\sim \\mathrm{keV} $ favors a {\\bf cored distribution}. Notice that a bosonic thermal relic that features a BEC can behave as CDM with small velocity dispersion and a primordial phase space density consistent with cusped distributions if $ T_d \\ll T_c $. However, these BEC DM candidates must decouple at a temperature scale {\\it higher} than the electroweak. Recent results from $N$-body simulations suggests that the phase space density relaxes by a factor $ \\sim 10^2 $ during gravitational clustering for $ 0 \\leq z \\leq 10 $ \\cite{numQ}. Combining these numerical results with the observational results on dSphs \\cite{gilmore} and the present DM density, we conclude that the mass of \\emph{thermal relics} that decoupled when ultrarelativistic is \\be \\label{masarango2} m_{cored}\\sim \\frac2{g^\\frac14} \\; \\mathrm{keV} \\quad , \\quad m_{cusp} \\sim\\frac8{g^\\frac14} \\; \\mathrm{keV} \\; . \\ee The decoupling temperature for the DM candidate that would favor cusped profiles must be near a grand unified scale for a large symmetry group with $g_d \\gtrsim 2000$ which effectively results in a colder relic today with a far smaller velocity dispersion. \\medskip The \\emph{enormous} discrepancy between the primordial phase space density for WIMPs of $ m\\sim 100\\,\\mathrm{GeV}; \\; T_d \\sim 10\\, \\mathrm{MeV} $, eq.(\\ref{wimpo}) and the phase space densities in dSphs, either cored (eq.\\ref{core2} ) or cusped (eq.\\ref{cusp2}) cannot be explained by the two orders of magnitude of gravitational relaxation of phase space densities found with recent $N$-body simulations \\cite{numQ}, although these initialize the simulation with much smaller values of the primordial phase space density. \\medskip We have studied a scenario for decoupling out of equilibrium motivated by previous studies of particle production and thermalization via an UV cascade. The distribution function obtained from previous studies \\cite{dvd}, remarkably describes the non-equilibrium distribution functions for sterile neutrinos produced either resonantly \\cite{este} or non-resonantly \\cite{dw} as well as a recently proposed model for halo structure \\cite{strigari}. Our bounds in terms of arbitrary distribution functions lead to the following bounds on the mass, phase space density and velocity dispersion of these light relics that decoupled out of LTE: \\begin{itemize} \\item{For sterile neutrinos produced non-resonantly via the Dodelson-Widrow mechanism \\cite{dw} we find \\be \\frac{1.04 \\; \\mathrm{keV}}{g^\\frac14} \\leq m \\leq \\frac{g_d}{g} \\; 46.5 \\, \\mathrm{eV} \\quad , \\quad \\frac{\\rho_{DM}}{\\sigma^3_{DM}} = 0.57 \\; g \\times 10^5 \\; \\Big[\\frac{m}{\\mathrm{keV}}\\Big]^3 \\; \\frac{M_\\odot/\\mathrm{kpc}^3}{\\big(\\mathrm{km}/\\mathrm{s}\\big)^3} \\quad , \\quad \\sigma_{DM} = \\frac{0.187 }{g^{\\frac13}_d} \\; \\Big(\\frac{\\mathrm{keV}}{m}\\Big) \\; \\Big(\\frac{\\mathrm{km}}{s}\\Big) \\; . \\label{dw2} \\ee The upper and lower bound on the mass can only be compatible if the sterile neutrino decouples with $g_d \\gtrsim 20-30$. For $m \\sim \\mathrm{keV}$ the primordial phase space density is compatible with cored but not with cusped profiles in the dShps data \\cite{gilmore}. Combining these bounds with the results from $N$-body simulations on the relaxation of the phase space density \\cite{numQ} and with the observational constraint eq.(\\ref{rss}) \\cite{gilmore}, we obtain the value \\be m\\sim \\frac{4}{g^{\\frac13}} \\; \\mathrm{keV} \\label{boundste2} \\ee for the mass of sterile neutrinos produced non-resonantly by the Dodelson-Widrow mechanism. } \\item{For sterile neutrinos produced by a net-lepton number driven resonant conversion \\cite{este} we find \\be \\frac{289\\,\\mathrm{eV}}{g^\\frac14} \\leq m \\leq \\frac{g_d}{g}\\; 81.4\\; \\mathrm{eV} \\quad , \\quad \\frac{\\rho_{DM}}{\\sigma^3_{DM}} = 9.6 \\; g \\times 10^6\\,\\Big[\\frac{m}{\\mathrm{keV}}\\Big]^4 \\; \\frac{M_\\odot/\\mathrm{kpc}^3}{\\big(\\mathrm{km}/\\mathrm{s}\\big)^3} \\quad , \\quad \\sigma_{DM} = \\frac{0.028 }{g^{\\frac13}_d} \\; \\Big(\\frac{\\mathrm{keV}}{m}\\Big) \\; \\Big(\\frac{\\mathrm{km}}{s}\\Big) \\; . \\label{sterQ2} \\ee The small velocity dispersion is a consequence of the distribution function being skewed towards small momentum. Again for $ m \\sim \\mathrm{keV} $, the primordial phase space density is compatible with cored but not cusped profiles in the dShps data \\cite{gilmore}. For sterile neutrinos produced by resonant conversion, a similar analysis as for the previous case yields \\be m\\sim \\frac{0.8}{g^\\frac14} \\; \\mathrm{keV} \\; . \\label{boundeste} \\ee } \\item{For the model proposed in ref. \\cite{strigari} we find \\be \\frac{475 \\; \\mathrm{eV}}{(\\beta \\; g)^\\frac14} \\leq m \\leq \\frac{g_d}{\\beta g} \\; 1.94 \\; \\mathrm{eV} \\quad , \\quad \\frac{\\rho_{DM}}{\\sigma^3_{DM}} = 1.33 \\; \\beta \\; g \\times 10^6 \\; \\Big[\\frac{m}{\\mathrm{keV}}\\Big]^4 \\; \\frac{M_\\odot/\\mathrm{kpc}^3}{\\big(\\mathrm{km}/\\mathrm{s}\\big)^3} \\quad , \\quad \\sigma_{DM} = \\frac{0.187}{g^{\\frac13}_d} \\; \\Big(\\frac{\\mathrm{keV}}{m}\\Big) \\; \\Big(\\frac{\\mathrm{km}}{s} \\Big)\\; . \\label{bounstri}\\ee } \\end{itemize} It is noteworthy that the $N$-body results of ref. \\cite{numQ} which yield phase space relaxation by a factor $ \\sim 10^2 $ bring the values of the primordial phase space density of the above cases within the range consistent with the phase space densities for cored profiles in dSphs \\cite{gilmore} for $ m \\sim \\mathrm{keV} $. On the contrary, in the case of WIMPs with $m\\sim 100\\,\\mathrm{GeV},T_d\\sim 10\\,\\mathrm{MeV}$, relaxation by {\\bf many} orders of magnitude is necessary for their phase space densities to be compatible with the observed values both for cores and for cusps. Therefore the bounds eqs.(\\ref{boundste2})-(\\ref{boundeste}) confirm that $ \\sim \\mathrm{keV} $ relics that decouple out of equilibrium while ultrarelativistic via the mechanisms described above yield values for phase space densities that are in agreement with cores in the DM distribution. \\medskip The results obtained in this article for the new mass bounds, primordial phase space densities and velocity dispersion in term of arbitrary, but homogeneous and isotropic distribution functions establish a link between the microphysics of decoupling and observable quantities. They also warrant deeper scrutiny of the non-equilibrium aspects of sterile neutrinos\\cite{boyho} for a firmer assessment of their potential as DM candidates." }, "0710/0710.4003_arXiv.txt": { "abstract": "The stellar evolution code YREC is outlined with emphasis on its applications to helio- and asteroseismology. The procedure for calculating calibrated solar and stellar models is described. Other features of the code such as a non-local treatment of convective core overshoot, and the implementation of a \\cws{parametrized} description of turbulence in stellar models, are considered in some detail. The code has been extensively used for other astrophysical applications, some of which are briefly mentioned at the end of the paper. ", "introduction": "The aim of this paper is to provide an overview of the Yale Rotating Stellar Evolution Code (YREC), as it has been applied in the last few years to research in helio- and asteroseismology. Although YREC contains extensions to model the effects of rotation in an oblate coordinate system, we describe here the ``non-rotating'' version. In addition to a general description, we shall emphasize three features of the code which have been implemented because of their special relevance to seismology. The first feature is the procedure utilized for the automatic calculation of calibrated solar and stellar models whose pulsational properties are to be investigated. The second feature is the treatment of convective core overshoot. Finally, the third feature is the implementation in stellar models of the effects of turbulence on the structure of the surface layers of stars with a convective envelope. The \\cws{parametrization} of turbulence to one dimension is based on three-dimensional radiative hydrodynamical (3D HRD) simulations of the highly superadiabatic layer (SAL) in the atmosphere. The interaction of turbulent convection and radiation in these thin transition regions is poorly known. Oscillation frequencies are sensitively affected by the structure of transition regions between radiative and convective layers. Seismology thus offers a unique opportunity to explore a long standing problem in stellar physics. Like most stellar evolution codes, YREC is a continuously evolving research tool to which many have contributed. As a result, different versions of YREC are in use at several institutions, which have been applied to a variety of research purposes. Some of the most significant applications of YREC are listed in the text and at the end of this paper (see Sect.~\\ref{other}). The rotating version of YREC, originally developed by \\citet{1988PhDT........14P}, includes a 1.5D treatment of rotation, extending the work of \\citet{1970stro.coll...20K} and \\citet{1981ApJ...243..625E}, and using the formalism of \\citet{1980PhDT.........5L}. A 2D version of YREC has also recently been implemented, specifically to address some fundamental aspects of solar magnetic activity \\citep{2006ApJS..164..215L}. Sect.~\\ref{num} outlines the numerical scheme adopted to solve the classical differential equations of stellar structure and evolution. The treatment of the boundary conditions, of special importance for seismology, are described in Sect.~\\ref{bc}. The constitutive physics, i.e. the equation of state and radiative and conductive opacities, are reviewed in Sect.~\\ref{eos}, and the nuclear processes are described in Sect.~\\ref{nuc}. Stellar physics topics such as superadiabatic convection, element diffusion, convective core overshoot, and turbulence in the outer layers, all of which also have important seismological signatures, are covered in Sect.~\\ref{sbc}, Sect.~\\ref{diff}, Sect.~\\ref{ov} and Sect.~\\ref{turb}, respectively. The operation of the code is described in Sect.~\\ref{runyrec}, with emphasis on helio- and asteroseismic applications. Seismic diagnostics applications are described in Sect.~\\ref{dia}. The role played by YREC in the research on solar neutrinos and helio-seismology is summarized in Sect.~\\ref{solar}. Studies of advanced evolutionary phases and applications to stellar population studies are listed in Sect.~\\ref{other}. ", "conclusions": "" }, "0710/0710.4145_arXiv.txt": { "abstract": "Precision infrared photometry from Spitzer has enabled the first direct studies of light from extrasolar planets, via observations at secondary eclipse in transiting systems. Current Spitzer results include the first longitudinal temperature map of an extrasolar planet, and the first spectra of their atmospheres. Spitzer has also measured a temperature and precise radius for the first transiting Neptune-sized exoplanet, and is beginning to make precise transit timing measurements to infer the existence of unseen low mass planets. The lack of stellar limb darkening in the infrared facilitates precise radius and transit timing measurements of transiting planets. Warm Spitzer will be capable of a precise radius measurement for Earth-sized planets transiting nearby M-dwarfs, thereby constraining their bulk composition. It will continue to measure thermal emission at secondary eclipse for transiting hot Jupiters, and be able to distinguish between planets having broad band emission {\\it vs.} absorption spectra. It will also be able to measure the orbital phase variation of thermal emission for close-in planets, even non-transiting planets, and these measurements will be of special interest for planets in eccentric orbits. Warm Spitzer will be a significant complement to Kepler, particularly as regards transit timing in the Kepler field. In addition to studying close-in planets, Warm Spitzer will have significant application in sensitive imaging searches for young planets at relatively large angular separations from their parent stars. ", "introduction": "The Spitzer Space Telescope (\\citet{Werner2004}) was the first facility to detect photons from known extrasolar planets (\\citet{Charbonneau2005, Deming2005}), inaugurating the current era wherein planets orbiting other stars are being studied directly. Cryogenic Spitzer has been a powerful facility for exoplanet characterization, using all three of its instruments. Spitzer studies have produced the first temperature map of an extrasolar planet (\\citet{Knutson2007a}), and the first spectra of their atmospheres (\\citet{Grillmair2007, Richardson2007}). Spitzer will continue to study exoplanets when its store of cryogen is exhausted. `Warm Spitzer' (commencing $\\sim$ spring 2009) will remain at T $\\sim$~35K (passively cooled by radiation), allowing imaging photometry at 3.6 and 4.5 $\\mu$m, at full sensitivity. The long observing times that are projected for the warm mission will facilitate several pioneering exoplanet studies not contemplated for the cryogenic mission. ", "conclusions": "" }, "0710/0710.0370_arXiv.txt": { "abstract": "This paper describes the MegaPipe image processing pipeline at the Canadian Astronomical Data Centre. The pipeline combines multiple images from the MegaCam mosaic camera on CFHT and combines them into a single output image. MegaPipe takes as input detrended MegaCam images and does a careful astrometric and photometric calibration on them. The calibrated images are then resampled and combined into image stacks. The astrometric calibration of the output images is accurate to within 0.15 arcseconds relative to external reference frames and 0.04 arcseconds internally. The photometric calibration is good to within 0.03 magnitudes. The stacked images and catalogues derived from these images are available through the CADC website. ", "introduction": "The biggest barrier to using archival MegaCam \\markcite{megacam}({Boulade} {et~al.} 2003) images is the effort required to process them. While the individual images are occasionally useful by themselves, more often the original scientific program called for multiple exposures on the same field in order to build up depth and get rid of image defects. Therefore, the images must be combined. A typical program calls for 5 or more exposures on a single field. Each MegaCam image is about 0.7Gb (in 16-bit integer format), making image retrieval over the web tedious. Because of the distortion of the MegaPrime focal plane, the images must be resampled. This involves substantial computational demands. During this processing, which is often done in a 32-bit format, copies of the images must be made, increasing the disk usage. In summary, the demands in terms of CPU and storage are non-trivial. Presumably Moore's law (1965), \\markcite{moore} will make these concerns negligible, if not laughable, in ten years time. However, at the moment, they present a technological barrier to easy use of MegaCam data. The Elixir pipeline \\markcite{elixir}({Magnier} \\& {Cuillandre} 2004) at CFHT processes each MegaCam image. It does a good job of detrending (bias-subtracting, flat-fielding, de-darking) the images. However, the astrometric solution Elixir provides is only good to 0.5-1.0 arcseconds. To combine the images, they must be aligned to better than a pixel. One arcsecond accuracy is insufficient. Therefore, it is necessary for a user to devise some way of aligning the images to higher accuracy. This is not an easy task, and rendered more difficult by the distortion of the MegaPrime focal plane. The problem is not intractable and there do exist a number of software solutions to the problem, but it remains an obstacle to easy use of MegaCam data. In short, while the barriers to using archival MegaCam data are not insurmountable, they make using these data considerably less attractive. MegaPipe aims to increase usage of MegaCam data by removing these barriers. This paper describes the MegaPipe image processing pipeline. MegaPipe combines MegaCam images into stacks and extracts catalogues. The procedure can be broken down into the following steps: \\begin{itemize} \\item Image grouping (Section \\ref{grouping}) \\item Astrometric calibration (Section \\ref{astrom}) \\item Photometric calibration (Section \\ref{photom}) \\item Image stacking (Section \\ref{comb}) \\item Catalogue generation (Section \\ref{cat}) \\end{itemize} Sections \\ref{qualastro} and \\ref{qualphoto} discuss checks on the astrometric and photometric properties of the output images. Section \\ref{sec:prod} describes the production and distribution of the images. ", "conclusions": "" }, "0710/0710.5591.txt": { "abstract": "We study X-ray spectra of Cyg X-3 from \\sax, taking into account absorption and emission in the strong stellar wind of its companion. We find the intrinsic X-ray spectra are well modelled by disc blackbody emission, its upscattering by hot electrons with a hybrid distribution, and by Compton reflection. These spectra are strongly modified by absorption and reprocessing in the stellar wind, which we model using the photoionization code {\\tt cloudy}. The form of the observed spectra implies the wind is composed of two phases. A hot tenuous plasma containing most of the wind mass is required to account for the observed features of very strongly ionized Fe. Small dense cool clumps filling $\\la$0.01 of the volume are required to absorb the soft X-ray excess, which is emitted by the hot phase but not present in the data. The total mass-loss rate is found to be (0.6--$1.6)\\times 10^{-5}\\msun$ yr$^{-1}$. We also discuss the feasibility of the continuum model dominated by Compton reflection, which we find to best describe our data. The intrinsic luminosities of our models suggest that the compact object is a black hole. ", "introduction": "Cyg X-3 is a high-mass X-ray binary (XRB) system with a short orbital period, $P=4.8$ h. It is located at a distance of $d\\sim 9$ kpc in the Galactic plane (Dickey 1983, assuming the 8 kpc distance to the Galactic Centre; Predehl et al.\\ 2000). Due to strong interstellar absorption, its optical counterpart is not observable, though infrared observations indicate it is a Wolf-Rayet (WR) star (van Kerkwijk et al.\\ 1996). In spite of its discovery already in 1966 (Giacconi et al.\\ 1967), it remains poorly understood. In particular, it remains uncertain whether its compact object is a black hole or a neutron star. Currently, there are two other known X-ray binaries containing WR stars, IC 10 X-1 (Prestwich et al.\\ 2007) and NGC 300 X-1 (Carpano et al.\\ 2007a, b). In both cases, there is dynamical evidence that the compact object is a black hole. Cyg X-3 is the brightest radio source among X-ray binaries (Mccollough et al.\\ 1999). It shows very strong radio outbursts and resolved jets (Marti, Paredes \\& Peracaula 2000; Mioduszewski et al.\\ 2001). In the X-rays, it exhibits a wide range of variability patterns. In particular, transitions between the hard and soft spectral state occur on the timescale of months to years. The understanding of the radiative processes underlying the X-ray spectra of Cyg X-3 remains rather rudimentary. Zdziarski \\& Gierlinski (2004) have shown an overall similarity of the spectral states of Cyg X-3 to canonical states of black hole binaries. However, details, in particular the energy of the cutoff in the hard state, appear to differ significantly. Many different models have been fitted to spectra of Cyg X-3 by various authors, but each of them appeared to fit the data well, which precluded determination of the correct one. The models assumed different system components, geometry, radiative processes and the absorbing medium. They included different combinations of absorbers, Gaussian lines, blackbody, power-law with or without a cutoff, and bremsstrahlung, and different model combinations were proposed for the hard and soft states, see, e.g., White \\& Holt (1982), Nakamura et al.\\ (1993), Rajeev et al.\\ (1994). Later, the power-law models were replaced by more physical models of Comptonization, see Vilhu et al.\\ (2003), Hjalmarsdotter et al.\\ (2004) and Hjalmarsdotter et al.\\ (2008, hereafter H08). The interpretation of the intrinsic unabsorbed spectra, remains, however, not clear. In particular, H08 show that {\\it INTEGRAL\\/} data of Cyg X-3 in the hard state could be well modelled by three models with very different shapes of the intrinsic continuum, yet the model most promising from a physical point of view yielded the worst statistical fit. A key issue in the search for understanding the spectra of Cyg X-3 appears to be the nature of the complex intrinsic absorption, most likely caused by the wind from the companion star. In this paper, we study the effects of the WR stellar wind on the X-ray spectra. In Section \\ref{data}, we present the data used for our study. Then, we discuss our theoretical models in Section \\ref{s:model}. In Section \\ref{results}, we describe the complex interactions between the X-rays and the wind, and present results of our fits to the data. Sections \\ref{discussion} and \\ref{conclusions} contain a discussion of our results and our conclusions, respectively. ", "conclusions": "\\label{conclusions} There are a number of essential features of our model which are new in the history of the studies of Cyg X-3. We have used a realistic model of the WR star, with a heavy small inner core and dense wind, and with the abundances typical to helium stars. We have used an accurate model of Comptonization together with accurate radiative transfer calculations to model the intrinsic X-ray emission in the vicinity of the compact object reprocessed by the surrounding wind. We have used high quality \\sax\\/ data, which are characterized by both relatively good spectral resolution and broad-band coverage. We have used the X-rays emission as a probe of the wind properties. We have determined the wind mass-loss rate to be (0.6--$1.6)\\times 10^{-5}\\msun$ yr$^{-1}$, which is consistent with mass-loss rates of WR stars. We have disproved a previous claim that if Cyg X-3 contains a WR star its wind would be completely opaque to X-rays. We have also determined the structure of the wind illuminated by the X-rays, and found it contains both a hot phase (containing most of its mass) and cold dense clumps, occupying less than 1 per cent of the wind volume. The hot phase was needed to account for the strong $\\sim$9 keV edge seen in the data, and the cold phase was needed to absorb a significant soft-excess emission produced by the hot phase. We have also found that the density enhancement in the clumps averaged over both phases is consistent with the clumping factor found in WR stellar winds using other methods. The lines emitted by the wind found in the \\sax\\/ data are consistent with the high-resolution measurement by {\\it Chandra}. Our fits imply that the observed X-ray continua contain very strong Compton reflection components. This would point out to the presence of a geometrically thick torus in the accretion flow, which would obscure most of the intrinsic X-ray source but allow us to see its reflection. The reflecting material is found to be weakly ionized (from the presence of an edge at $\\sim$7 keV), but no associated fluorescent Fe K$\\alpha$ (at 6.4 keV) is found. This presents us with a puzzle remarkably similar to that encounted in a number of Narrow-Line Seyfert 1 galaxies. On the other hand, the finding of the dominance of reflection may be an artifact of not including enough details in our spectral modelling. In particular, an additional partial covering by a weakly ionized medium could be responsible for the $\\sim$7-keV edge in the data. Based on the bolometric luminosity of the inferred intrinsic continua, the compact object in Cyg X-3 most likely a black hole." }, "0710/0710.4623_arXiv.txt": { "abstract": "{ We present the analysis and results of 12.5 hours of high-energy gamma-ray observations of the EGRET-detected pulsar PSR B1951+32 using the Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE). STACEE is an atmospheric Cherenkov detector, in Albuquerque, New Mexico, that detects cosmic gamma rays using the shower-front-sampling technique. STACEE's sensitivity to astrophysical sources at energies around 100 GeV allows it to investigate emission from gamma-ray pulsars with expected pulsed emission cutoffs below 100 GeV. We discuss the observations and analysis of STACEE's PSR 1951+32 data, accumulated during the 2005 and 2006 observing seasons. } \\begin{document} ", "introduction": "Young energetic pulsars (rapidly rotating, highly magnetized neutron stars, that produce non-thermal photons) remain as yet the only Galactic objects unambiguously detected within the EGRET energy range. To date, more than 1500 pulsars have been observed at radio energies \\cite{manchester05}. About 70 of these are seen in X-rays \\cite{kaspi06} but only 6 (Crab, Geminga, Vela, PSR B1951+32, PSR1706-44, and PSR B1055-52) were detected by EGRET \\cite{nolan96}. Of these, only one, PSR B1951+32, has been observed up to $\\sim$20~GeV without any apparent cutoff in its differential energy spectrum \\cite{ramanamurthy95}. Accordingly, PSR 1951+32 is often considered the best pulsar candidate for pulsed-emission detection in the very-high-energy (VHE) regime; i.e.\\ at energies above $\\sim$50~GeV. As is the case for most pulsars, PSR B1951+32 was first detected at radio energies. It was discovered with a 39.5~ms period in the radio synchrotron nebula CTB 80 \\cite{kulkarni98}. From the radio observations it was deduced that the pulsar has a characteristic age of 1.1$\\times$10$^5$~yr with a surface magnetic field of 4.9$\\times$10$^{11}$~G, and a rotational energy loss rate of 3.7$\\times$10$^{36}$~ergs~s$^{-1}$. It has also been observed in X-rays. While the pulsar emits a single pulse at radio frequencies and in X-rays, the EGRET observations display a double-pulsed profile with neither of the gamma-ray peaks coinciding with the radio peak. Interestingly, there is no evidence for any interpeak emission in the gamma-ray data \\cite{ramanamurthy95}. At TeV energies, only upper limits on the pulsed emission, and on the emission from the surrounding synchrotron nebula, exist \\cite{srinivasan97}. A recent report by the MAGIC collaboration \\cite{albert07} presents observations above 75~GeV with no evidence for pulsed emission. % Models that attempt to explain the non-thermal high-energy pulsed emission from pulsars generally fall into one of two broad categories; the Polar Cap \\cite{muslimov03} or Outer Gap \\cite{hirotani01} models. While both models can explain the observed gamma-ray emission at EGRET energies, they differ in their predictions for detectable emission above 20~GeV. In both scenarios the pulsed emission is attributed to particle acceleration in the pulsar's magnetosphere, with spectral cutoffs predicted in the 20-100~GeV energy range. The Polar Cap model localizes the emission site to a region close to the magnetic poles of the neutron star, where the magnetic field is strong. The Outer Gap model, on the other hand, contends that gamma-ray production occurs far from the neutron star surface in a region of relatively weak magnetic field, in so called ``outer gaps'' near the null surface of the outer magnetosphere. The emission sites dictate the energy of the expected spectral cutoff, insofar as the maximum energy of the curvature-radiated photons escaping the magnetosphere is limited by pair-production in the pulsar's magnetic field. Since the magnetic field is stronger near the polar cap, the Polar Cap model anticipates a lower-energy cutoff than the Outer Gap model. Any detection of TeV emission would clearly favor the Outer Gap model and would thereby significantly contribute to our understanding of pulsar emission processes. Thus far, no pulsed TeV emission has been detected from any pulsar, although ever-lower upper limits are constraining the emission models. As a gamma-ray observatory operating at the lower end of the VHE regime (around 100 GeV) the Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE) is suited to pulsar observations. STACEE observations of PSR B1951+32, undertaken in 2005 and 2006, are reported here. ", "conclusions": "No evidence for pulsed gamma-ray emission above 117 GeV was found in the selected gamma-ray data set used in this work (Figure 1). To calculate flux upper limits, we used the method of Helene \\cite{helene83} to estimate 99.0\\% upper limits for excess events within the pulsed phase profile seen by EGRET at lower energies \\cite{ramanamurthy95}. That is, emission is assumed to occur in the phase range of the main pulse, phase 0.12--0.22, and the interpulse, phase 0.48--0.74. A differential flux upper limit of 4.52~$\\times 10^{-6}$~MeV~cm$^{-2}$~s$^{-1}$ was determined at the energy threshold of 117 GeV, by extrapolation of the EGRET spectrum to STACEE energies. \\bigskip {\\bf Acknowledgements:} Many thanks go to the staff of the National Solar Tower Test Facility, who have made this work possible. This work was funded in part by the US National Science Foundation, the Natural Sciences and Engineering Research Council of Canada, Fonds Quebecois de la Recherche sur la Nature et les Technologies, the Research Corporation, and the University of California at Los Angeles." }, "0710/0710.0402.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory \\abstract % context heading (optional) % {} leave it empty if necessary % {The cosmic X-ray background is the summed radiation from the growth of the massive black holes that reside in the centres % of galaxies. By studying the population of sources that produce the X-ray background through X-ray surveys at various depths, % we can determine when, where and how this growth occurs. } % aims heading (mandatory) {} {X-ray sources at intermediate fluxes (a few $\\times 10^{-14}\\, {\\rm erg\\, cm^{-2}\\, s^{-1}}$) with sky density of $\\sim 100\\, {\\rm deg}^{-2}$, are responsible for a significant fraction of the cosmic X-ray background at various energies below 10 keV. The aim of this paper is to provide an unbiased and quantitative description of the X-ray source population at these fluxes and in various X-ray energy bands.} % methods heading (mandatory) {We present the XMM-Newton Medium sensitivity Survey (XMS), including a total of 318 X-ray sources found among the serendipitous content of 25 XMM-Newton target fields. The XMS comprises four largely overlapping source samples selected at soft (0.5-2 keV), intermediate (0.5-4.5 keV), hard (2-10 keV) and ultra-hard (4.5-7.5 keV) bands, the first three of them being flux-limited. } % results heading (mandatory) {We report on the optical identification of the XMS samples, complete to 85-95\\%. At the flux levels sampled by the XMS we find that the X-ray sky is largely dominated by Active Galactic Nuclei. The fraction of stars in soft X-ray selected samples is below 10\\%, and only a few per cent for hard selected samples. We find that the fraction of optically obscured objects in the AGN population stays constant at around 15-20\\% for soft and intermediate band selected X-ray sources, over 2 decades of flux. The fraction of obscured objects amongst the AGN population is larger ($\\sim 35-45\\%$) in the hard or ultra-hard selected samples, and constant across a similarly wide flux range. The distribution in X-ray-to-optical flux ratio is a strong function of the selection band, with a larger fraction of sources with high values in hard selected samples. Sources with X-ray-to-optical flux ratios in excess of 10 are dominated by obscured AGN, but with a significant contribution from unobscured AGN. } % conclusions heading (optional), leave it empty if necessary {} ", "introduction": "Supermassive black holes (SMBHs, i.e., with masses $\\sim 10^6-10^9\\, {\\rm M}_{\\odot}$) have been detected in the centers of virtually all nearby galaxies \\citep{Merrit01,Tremaine02}. In many of these galaxies -including our own-, the SMBH is largely dormant, i.e., the luminosity is many orders of magnitude below the Eddington limit. Only $\\sim 10\\%$ of today's galaxies (at most) host active galactic nuclei (AGN), and a very large fraction of them are in fact inconspicuous at most wavelengths because of obscuration \\citep{Fabian99}. %SMBH %masses correlate well with host galaxy properties (stellar or gas %central velocity dispersion, bulge luminosity), and are estimated %to contain around 0.4-0.6\\% of the bulge masses %\\citep{Magorrian98}. It is generally believed that the seeds of these SMBHs were the remnants of the first generation of massive stars in the history of the Universe. These early black holes may have had masses of tens of ${\\rm M}_{\\odot}$ at most. The growth of these relic black holes to their current sizes is very likely dominated by accretion, with additional contributions by other phenomena like black hole mergers and tidal capture of stars \\citep{Marconi04}. According to current synthesis models, the integrated X-ray emission produced by the growth of SMBHs by accretion over the history of the Universe is recorded in the X-ray background (XRB). Thus the XRB can be used to constrain the epochs and environments in which SMBHs developed. There are currently a number of existing or on-going surveys in various X-ray energy bands (see \\citet{Brandt05} for a recent compilation). In the pre-Chandra and pre-{\\it XMM-Newton} era the Einstein Extended Medium Sensitivity Survey \\citep{Maccacaro82,Gioia90,Stocke91} pioneered the procedure of determining typical X-ray to optical flux ratios for different classes of X-ray sources to facilitate the identification processes and has set the standards for serendipitous X-ray surveys. $ROSAT$ produced a number of surveys in the soft 0.5-2 keV X-ray band at various depths, e.g., the $ROSAT$ Bright Survey \\citep{Schwope00}, the intermediate flux RIXOS survey \\citep{Mason00} and the $ROSAT$ deep surveys \\citep{McHardy98,Georgantopoulos96,Hasinger98,Lehmann01} among others. These surveys show that AGN dominate the high Galactic latitude soft X-ray sky at virtually all relevant fluxes. The majority of these AGN are of spectroscopic type 1, which means that we are witnessing the growth of SMBH through unobscured lines of sight. In a moderate fraction of the sources identified, however, there is evidence for obscuration as their optical spectra lack broad emission lines (type 2 AGN). \\citet{Ueda03} discuss the results from a large area X-ray survey in the 2-10 keV band with ASCA and those from HEAO-1 and $Chandra$, where a larger fraction of the sources identified correspond to type 2 AGN. With {\\it Chandra} and {\\it XMM-Newton} coming into operation X-ray surveys, particularly at energies above a few keV, have been significantly boosted. Thanks to the high sensitivity and large field of view of the EPIC cameras \\citep{Turner01,Struder01} on board {\\it XMM-Newton} \\citep{Jansen01}, X-ray surveys requiring large solid angles have been dominated by this instrument. The Bright Source Survey-BSS \\citep{Dellaceca04} contains 400 sources brighter than $\\sim 7\\times 10^{-14}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$ either in 0.5-4.5 keV or 4.5-7.5 keV. The BSS samples\\footnote{{\\tt http://www.brera.mi.astro.it/\\~{}xmm/}}, which have been identified to $\\sim 90\\%$ \\citep{Caccianiga07}, show an X-ray sky dominated by AGN, where the fraction of obscured objects varies with the selection band (sample selection at harder energies reveals a higher fraction of obscured objects as expected). Deep surveys have also been conducted by {\\it XMM-Newton}, for example in the Lockman Hole down to $\\sim 10^{-15} \\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$ \\citep{Hasinger01,Mateos05b}. However, thanks to its much better angular resolution, the {\\it Chandra} deep surveys are photon counting limited and far from confusion and are consequently much more competitive at fainter fluxes \\citep{Alexander03,Tozzi06}. Optical identification of these deep surveys is largely incomplete, a fact that is driven by the intrinsic faintness and red colour of most of the counterparts to the faintest X-ray sources. In the intermediate flux regime, however, the identified fractions are large and nearing completion. It is interesting to note that deep surveys start to find a population of galaxies not necessarily hosting active nuclei as an important ingredient. In addition, the AGN population is found to contain an important fraction of obscured objects. The wide range of intermediate X-ray fluxes, between say $10^{-15}\\, \\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$ and $10^{-13}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$ have also been the subject of a number of on-going surveys. Besides bridging the gap between wide and deep surveys, intermediate fluxes sample the region around the break in the X-ray source counts \\citep{Carrera07}, and therefore their sources are responsible for a large fraction of the X-ray background. Among these, we highlight the {\\it XMM-Newton} survey in the well-studied (at many bands) COSMOS field, which covers $2\\, \\deg^2$ to fluxes $\\sim 10^{-15}\\, {\\rm erg\\, cm^{-2}\\, s^{-1}}$ \\citep{Hasinger07}. The optical identification is still on-going, reaching 40\\% \\citep{Brusa07}. At fluxes around $10^{-14}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$, the HELLAS2XMM survey \\citep{Baldi02,Fiore03}, now extended to cover $1.4\\, \\deg^2$, contains over 220 X-ray sources, optically identified to 70\\% completeness \\citep{Cocchia07}. Other surveys in this flux range include the {\\it XMM-Newton} survey in the Marano field \\citep{Krumpe07}, which is 65\\% identified over a modest solid angle of $0.28\\, \\deg^2$. Also the XMM-2dF survey (Tedds et al., in preparation), which contains almost 1000 X-ray sources optically identified in the Southern Hemisphere, is an important contributor in this regime. {\\it Chandra} has also triggered surveys at intermediate fluxes, most notably the {\\it Chandra} Multiwavelength Survey \\citep{Kim04a,Kim04b,Green04}, covering $1.7\\, \\deg^2$ and identified to $\\sim 40\\%$ completeness \\citep{Silverman05}. In the realm of this variety of X-ray surveys that yield a qualitative picture of the X-ray sky, the {\\it XMM-Newton} Medium sensitivity Survey (XMS) discussed in this paper, finds its role in three important ways: a) it deals with very large samples, selected at various X-ray bands where {\\it XMM-Newton} is sensitive, from 0.5 to 10 keV; b) the samples that we consider have been identified almost in full, from 85\\% to 95\\% completeness and c) three out of the four samples that we explore are strictly flux limited in three energy bands (0.5-2 keV, 0.5-4.5 keV and 2-10 keV). Armed with these unique features, the XMS is a very powerful tool to derive a {\\it quantitative} characterization of the population of X-ray sources selected in various bands, and also to study and characterize minority populations, all of it at specific intermediate X-ray fluxes where a substantial fraction of the X-ray background below 10 keV is generated. The power of the XMS is enhanced by the fact that to some extent it is a representative sub-sample of the {\\it XMM-Newton} X-ray source catalogue 2XMM\\footnote{Pre-release under {\\tt http://xmm.vilspa.esa.es/xsa}}, containing 150,000 entries. Specific goals that have driven the construction of the XMS whose results are presented in this paper include: a) quantify the fraction of stars versus extragalactic sources at intermediate X-ray fluxes and at different X-ray energy bands; b) quantify the fraction of AGN that are classified as obscured by optical spectroscopy at intermediate X-ray fluxes and for samples selected in different energy bands; c) find the redshift distribution for the various classes of extragalactic sources and compare soft and hard X-ray selected samples; d) study the distribution of the X-ray-to-optical flux ratio for the various classes of X-ray sources, also as a function of X-ray selection band. The X-ray spectral properties of the sources of the XMS were already discussed in \\citet{Mateos05a}. Further goals that we will achieve with the XMS in forthcoming papers include: e) determine the fraction of ``red QSOs'' at intermediate X-ray fluxes and as a function of X-ray selection band; f) relate X-ray spectral properties (like photoelectric absorption) to optical colours of the counterpart; g) quantify the fraction of radio-loud AGN in the samples selected at various X-ray energies; h) construct Spectral Energy Distributions for the various classes of sources in the XMS. Results on these further aspects will be presented in a forthcoming paper (Bussons-Gordo et al., in preparation). The paper is organized as follows: in Section~\\ref{sec:XMS} we define the XMS along with the 4 samples that constitute it, including the X-ray source list; in Section~\\ref{sec:imaging} we discuss the multi-band optical imaging conducted on the {\\it XMM-Newton} target fields and the process for selecting candidate counterparts; this is continued in Section~\\ref{sec:identification} where we discuss the identification of the XMS sources in terms of optical spectroscopy, and list photometric and spectroscopic information on each XMS source. Section~\\ref{sec:XMSpopulations} presents the first scientific results from the XMS, specifically a description of the overall source populations, the fraction of stars in the various samples, the fraction of optically obscured AGN, and the X-ray to optical flux ratio of the different source populations. Section~\\ref{sec:conclusions} summarizes our main results. To clarify the terminology used in this paper, an AGN not displaying broad emission lines in its optical spectrum is termed as type 2 or obscured, and type 1 or unobscured otherwise. The property of being absorbed or unabsorbed refers only to the detection or not of photoelectric X-ray absorption. Throughout this paper, we used a single power law X-ray spectrum to convert from X-ray source count rate to flux in physical units, with a photon spectral index $\\Gamma=1.8$ for the XMS-S and XMS-X samples and $\\Gamma=1.7$ for the XMS-H and XMS-U samples. These are the average values obtained by \\citet{Carrera07}, which -as opposed to what we do here- used the specific value of $\\Gamma$ for each individual source and energy range. When computing luminosities, we also use the above spectra for K-correction and the concordance cosmology parameter values: $H_0=70\\, {\\rm km}\\, {\\rm s}^{-1}\\, {\\rm Mpc}^{-1}$, $\\Omega_m=0.3$ and $\\Omega_\\Lambda=0.7$. All quoted uncertainties in parameter estimates are shown at 90\\% confidence level for one interesting parameter. ", "conclusions": "\\label{sec:conclusions} In this paper we have presented the {\\it XMM-Newton} Medium sensitivity Survey XMS, and extracted a number of robust quantitative conclusions about the population of high Galactic latitude X-ray sources at intermediate flux levels. We have argued that given the completeness of our identifications and the relatively large size of the XMS samples, these conclusions can be safely exported to a much larger X-ray source catalogue like 2XMM. Our conclusions can be summarized as follows: \\begin{enumerate} \\item The high galactic latitude X-ray sky at intermediate flux levels is dominated by AGN, which includes type-1 and type-2 AGN as well as the so-called XBONG which are likely to host a low luminosity or obscured nucleus (or both). The stellar content is less than 10\\% in soft X-ray selected samples, and drops to below 5\\% at around soft X-ray fluxes $\\sim 10^{-14}\\, {\\rm erg}\\, {\\rm cm}\\, {\\rm s}^{-1}$. The stellar content in hard X-ray selected samples does not exceed a few per cent at most. Selection in 0.5-4.5 keV produces intermediate results. \\item Given the limited sensitivity of {\\it XMM-Newton} above a few keV -which is due to the roll over of effective area- current surveys conducted in the so-called ultra-hard band (4.5-7.5 keV) do not bring any new source population or any significant difference with respect to 2-10 keV selected surveys. Much longer exposure times would be needed to unveil any new heavily obscured population with {\\it XMM-Newton}. \\item Obscured AGN represent $\\sim 20\\%$ of the soft X-ray selected population of AGN, all the way from $\\sim 10^{-13}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$ down to $\\sim 10^{-15}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$, with no compelling evidence for an increase of this fraction towards fainter fluxes within this range. \\item Likewise, obscured AGN represent $\\sim 35\\%$ ($45\\%$ if all unidentified sources are obscured AGN) of the hard X-ray selected population of AGN, with no hint of an increase down to a hard X-ray flux $\\sim 10^{-14}\\, {\\rm erg}\\, {\\rm cm}^{-2}\\, {\\rm s}^{-1}$. \\item The fraction of X-ray sources with X-ray to optical flux ratio $>10$ (or $X/O>1$ using the notation of this paper) is a mere 3\\% in soft X-ray selected samples, but grows to 20\\% in hard X-ray selected samples. \\item Those sources with $X/O>1$ are mostly obscured AGN, but a fraction of around 20\\% of them in the hard band are unobscured type-1 AGN. This means that $X/O>1$ alone cannot be used as a proxy for obscured X-ray sources. \\end{enumerate}" }, "0710/0710.3659_arXiv.txt": { "abstract": "}{} \\textheight 240mm \\textwidth 170mm \\hoffset= 0mm \\voffset=0cm \\topmargin -2cm \\oddsidemargin 0mm \\evensidemargin 0mm \\begin{document} \\fontsize{12}{12} \\selectfont \\title{Luminosity Dependence of the Quasar Clustering from SDSS DR5} \\author{Ganna Ivashchenko} \\date{} 43024 objects, which were primarily identified as quasars in SDSS DR5 and have spectroscopic redshifts were used to study the luminosity dependence of the quasar clustering with the help of two different techniques. The obtained results reveal that brighter quasars are more clustered, but this dependence is weak, which is in agreement with the results by Porciani \\& Norberg \\cite{porciani2006} and theoretical predictions by Lidz et al. \\cite{lidz}. \\textbf{Key words:} quasars, large-scale structure, clustering, luminosity. PACS numbers: 98.65.Aj, 98.54.-h, 98.62.Ve. ", "introduction": "\\indent \\indent Determining the distribution of extragalactic objects is one of the most important problems of modern cosmology, because they are the only tracers of the dark matter, except anisotropy of the CMB, that helps us to realize the matter distribution on the redshifts of about 1000. Unfortunately even in the local Universe it is difficult to see the most faint galaxies, and when we go farther to larger redshifts, the most part of objects we can observe there with the modern ground-based telescopes used for large surveys are quasars as the most luminous objects. The matter distribution that we could reconstruct with the help of quasars is just like a light sketch, but it is all we have for today. The quasars are nonuniform objects: they have various luminosities and are on the different stages of their evolution. Thus the reasonable question arises: how their clustering could depend on their physical properties, for example on their luminosity? According to different numerical simulations of galaxy mergers that incorporate black hole growth, this dependence has to exist because more luminous quasars are considered to be born in denser environment. In the main part of such models, in which the host halo mass correlates with the instantaneous luminosity of the quasars (see e.g. \\cite{kaufman}), there should be a strong luminosity dependence of the quasar clustering. In the other type of such models, in which the host halo mass correlates with the peak luminosity of quasars, the luminosity dependence of the quasar clustering should be weaker because all of the quasars we see now, are considered to be similar objects but on different stages of their evolution (see \\cite{lidz} and references therein). It is worth to note, that the redshift distribution of the quasars is not the same for different luminosities due to possible evolution effects. The largest quasar surveys are 2dF and SDSS. In contrast to 2dF, which is finished for today and has 2QZ catalogue as a result \\cite{2dF-XII}, SDSS \\cite{SDSS5} is in progress and the area covered by it is increasing. Only a part of objects primarily classified as quasars were justified by spectra analysis and were included into SDSS Quasar Catalogue IV \\cite{dr4}. However even photometric classification of quasars with color diagrams \\cite{phot} is sufficient for using these objects for statistical purposes \\cite{myers2, myers1}. Some attempts to find luminosity dependence of the quasar clustering have been made with different samples. E.g. Adelberg \\& Steidel \\cite{adelberg}, who worked with their own survey, pointed to luminosity independent quasar clustering. 2dF-team, that studied 2QZ survey and found the little redshift evolution in the amplitude of the power spectrum \\cite{2dF-IV}, \\cite{2dF-XI} and significant increase in clustering amplitude at high redshifts \\cite{2dF-XIV, 2dF-II}, detected only marginal evidence for quasars with brighter apparent magnitudes to have a stronger clustering amplitude \\cite{2dF-IX}. Porciani and Norberg \\cite{porciani2006} also found weak luminosity dependence of the clustering in 2QZ survey. Furthermore they noted that samples with different redshifts show different trends in luminosity dependence: in the redshift bin $1.7$2. The host galaxy is surrounded by a giant Ly$\\alpha$ halo and the high rotation measure of the radio source means that it is embedded in a dense medium with an ordered magnetic field \\citep{Pentericci1997}. The radio galaxy is associated with a $>$3\\,Mpc-sized structure of galaxies with an estimated mass $>4\\times10^{14}$\\Msun\\ \\citep{Venemans2007}, indicating that it is the predecessor of a local rich cluster. Mechanical feedback from the active galactic nucleus (AGN) is sufficient to expel significant fractions of the interstellar medium of the massive gas-rich galaxy \\citep{Nesvadba2006}. Therefore the AGN may quench the star formation and allow the radio galaxy to evolve onto the red sequence as suggested by popular models of massive galaxy formation \\citep{Croton2006}. The study of the formation of the most massive galaxies through these and other mechanisms has so far mainly been limited to models and simulations \\cite[e.g][]{Dubinski1998,Gao2004,DeLucia2007}. Our aim is to use these observations of a high-redshift radio galaxy to study brightest cluster galaxy formation. In this work we present detailed images of widespread UV intergalactic light (IGL) situated in a 60\\,kpc halo that lies between the radio galaxy and the UV bright satellite galaxies. We examine the possible origins of this light and present in-situ star formation as the most probable. In section \\ref{method} we describe the observations and data reduction, in section \\ref{results} we describe the distribution and colour of the IGL, and discuss possible origins in section \\ref{nature}. Section \\ref{discussion} discusses the implications of widespread star formation in a halo around a massive forming brightest cluster galaxy. Throughout this work we use $H_0=71$, $\\Omega_M=0.27$, and $\\Omega_\\Lambda=0.73$ \\citep{Spergel2003}. All magnitudes are AB magnitudes. At a redshift of 2.156, the linear scale is 8.4\\,kpc/\". ", "conclusions": "The radio galaxy MRC\\,1138-262 is surrounded by a halo of IGL that extends across 60\\,kpc. After examining nebular continuum emission, synchrotron, inverse Compton, synchrotron self-Compton emission, scattering, and stripping of stars as possible sources of the IGL, we conclude the most likely origin of the IGL is in-situ star formation. The minimum star formation rate (i.e. uncorrected for dust) of the IGL is 57$\\pm8$\\Msunpyr, which is comparable to the total star formation rate derived from the UV luminosity integrated over all the galaxies within 70 kpc of the radio galaxy (and including the radio galaxy itself). Applying a minimum dust correction of $E(B-V)\\sim0.1$ imply by the red colours of the the IGL increases the star formation rate of the IGL to 142\\Msunpyr, and that of the whole Spiderweb system to more than 325\\Msunpyr. We estimate that the total observed star formation rate can produce approximately $\\sim$7\\% of the \\lya\\ emission in the halo surrounding the galaxy. The radial colour gradient of the IGL indicates that there is a smooth range in extinction or stellar population age from the outer to the inner parts of the halo, where the inner parts are older or dustier than the outer region. While the presence of large quantities of ionized gas found around several remarkable species of high redshift galaxies has been attributed to a number of proposed mechanisms that include AGN feedback, infall, cooling flows, starburst superwinds and mergers, these observations of MRC\\,1138-262 show that any successful model should be able to accommodate the mode of extended star formation present in this paper. Our data suggest that the formation of the most massive galaxies is connected with that of their gaseous envelopes and star forming halos, part of which may precede the intracluster light or cD envelopes, or perhaps may contribute to a satellite population. A significant amount of star formation might be occurring in the form of extended low surface brightness features, beyond the typical UV detection limits, as well as in largely obscured extended halos as detected at infrared and sub-mm wavelengths." }, "0710/0710.0589.txt": { "abstract": "Large-scale asymmetries in the stellar mass distribution in galaxies are believed to trace non-equilibrium situations in the luminous and/or dark matter component. These may arise in the aftermath of events like mergers, accretion, and tidal interactions. These events are key in the evolution of galaxies. In this paper we quantify the large-scale lopsidedness of light distributions in 25155 galaxies at $z < 0.06$ from the Sloan Digital Sky Survey Data Release 4 using the $m = 1$ azimuthal Fourier mode. We show that the lopsided distribution of light is primarily due to a corresponding lopsidedness in the stellar mass distribution. Observational effects, such as seeing, Poisson noise, and inclination, introduce only small errors in lopsidedness for the majority of this sample. We find that lopsidedness correlates strongly with other basic galaxy structural parameters: galaxies with low concentration, stellar mass, and stellar surface mass density tend to be lopsided, while galaxies with high concentration, mass, and density are not. We find that the strongest and most fundamental relationship between lopsidedness and the other structural parameters is with the surface mass density. We also find, in agreement with previous studies, that lopsidedness tends to increase with radius. Both these results may be understood as a consequence of several factors. The outer regions of galaxies and low-density galaxies are more susceptible to tidal perturbations, and they also have longer dynamical times (so lopsidedness will last longer). They are also more likely to be affected by any underlying asymmetries in the dark matter halo. ", "introduction": "%\\citet{rz95}: 18 galaxies. Galaxy disks exhibit a wide variety of shapes visible in the near-IR. They are due to distortions in surface density distributions, not mass-to-light ratios. Stellar velocities in nonaxisymmetric galaxies differ by 3-6\\% from those in symmetric ones. %\\citet{zr97}: 60 galaxies, magnitude-limited. 30\\% of galaxies have $A_1 > 0.2$. Lopsided mass distributions remain long enough to indicate past interactions, not just ongoing ones. Often the companion is not obvious, either merged or fled. %\\citet{rr98}: 54 early-type disk galaxies. Low SFRs to reduce possibility of asymmetric stellar distributions and increase possibilty of light->mass asymmetry tracing. Traces mass dependence because VRI bands show similar lopsidedness and it's OK to go shorter than I and K bands. 20\\% of galaxies have lopsidedness above 0.19. %\\citet{rrk00}: Correlations between lopsidedness and recent SF, and current SF were found, lopsided = star-forming. Significant fractions of stellar content can be created in a short time from minor mergers and on a similar timescale (1 Gyr). It has long been recognized that galaxies show large-scale asymmetries in their structure \\citep{bl+80}. Lopsided galaxies have such asymmetries where one side of their disk is more massive and/or more extended than the opposite side. This \u0093lopsidedness\u0094 can be traced in the spatial structure of the stars \\citep{rz95} and/or the HI gas \\citep{rs94} and/or in the large-scale kinematics of this material \\citep{ss+99}. There are a variety of mechanisms or events that have been proposed to produce the observed lopsidedness. All of them involve a time-dependent non-equilibrium dynamical state, in most cases triggered through an external process. Such external processes are a natural consequence of the standard Lambda Cold Dark Matter cosmological framework. This implies that galaxies assemble hierarchically (a process that is on-going). Examples that can lead to lopsidedness include a minor merger (\\citealt{wm+96}; \\citealt{zr97}), the tidal interaction resulting from a close encounter between roughly equal-mass galaxies \\citep{kl+02}, and the asymmetric accretion of intergalactic gas into the disk (\\citealt{bc+05}; \\citealt{kk+05}). Other mechanisms involve the dark matter halo: stars and gas orbiting in a lopsided dark matter halo (\\citealt{w94}; \\citealt{j97}; \\citealt{j99}) or a stellar/gas disk that is offset with respect to the center of the dark matter halo (\\citealt{ls98}; \\citealt{ns+01}). These also involve past tidal interactions and/or mergers that have perturbed the dark matter halo, but such perturbations may be quite long-lived. Finally, dynamical processes internal to the disk that lead to mildly lopsided distributions have also been investigated (\\citealt{st+90}; \\citealt{st96}; \\citealt{mt97}). A variety of programs to study lopsidedness have been undertaken over the past decade. Most of these investigations have studied the lopsided distribution of the stellar component through analysis of optical and near-infrared images. \\citet{zr97} studied a magnitude-limited sample of 60 field spiral galaxies. They measured lopsidedness as the radially averaged, azimuthal $m=1$ Fourier amplitude $A_1$ of the light (see Section \\ref{sec:error} below) and computed lopsidedness between 1.5 and 2.5 scale lengths in the galactic disks. The value of $A_1$ indicates the typical large-scale variation in mass density from side to opposite side at the same distance from the galactic center. The mass density typically varies from between $1 \\pm A_1$ times the average density at the same radius. They found that $\\sim 30$\\% of field spiral galaxies exhibited significant lopsidedness ($A_1 > 0.2$). \\citet{rr98} followed up this work by studying lopsidedness in 54 early-type galaxies and found that $\\sim 20$\\% had $A_1 > 0.19$. \\citet{cbj00} studied a sample of 113 $z < 0.01$ galaxies (elliptical, spiral, and irregular) but used a 180-degree rotational asymmetry measure $A_{180}$. They found that asymmetry is strongly dependent on morphological type, with lower asymmetry in elliptical and lenticular galaxies and higher asymmetry in late-type disk and irregular galaxies. More recently, \\citet{bc+05} have measured the Fourier $A_1$ parameter for 149 galaxies in the Ohio State University Bright Galaxy Survey. They confirmed that a large fraction of galaxies have significant lopsidedness in their stellar disks, with late-type galaxies being more lopsided. Lopsidedness in the {\\it light} distribution can be produced by either a corresponding asymmetry in the underlying {\\it mass} distribution in the stellar population or by large-scale variations in the mass-to-light ratio (e.g., from star formation and dust obscuration). \\citet{rz95} investigated this issue with a sample of 18 face-on spiral galaxies imaged in the K$^\\prime$ (2.2$\\mu$ m) band where the effect of young stars or dust is minimized. They found that about a third of the sample showed significant lopsidedness (similar to results from optical investigations). Similarly, \\citet{rr98} found that lopsidedness in early-type disk galaxies is nearly identical when observed in the $V$, $R$, and $I$ bands. They concluded that an asymmetric mass distribution then accounts for the majority of the asymmetry in the light distribution in these galaxies. Lopsidedness has also been studied in the distribution of HI gas. Since the HI can frequently be traced to significantly larger radii than the stars, these investigations are highly complementary to the optical image analysis. Due to the time-consuming nature of HI interferometric mapping, only modest size samples have been analyzed in this way \\citep{ss+99}. On the other hand, \\citet{rs94} have examined the global HI line profiles for roughly 1700 galaxies, and shown that at least 50\\% are significantly asymmetric (confirming that the large-scale HI distribution is frequently lopsided). HI maps also show that \u0096 apart from a lopsided distribution of the gas \u0096 the HI rotation curves are often asymmetric \\citep{ss+99}. The connection between the phenomena of structural and kinematic lopsidedness in galaxies is not yet clear \\citep{ss+99}. Despite these diverse investigations and the abundance of proposed models, the origin of lopsidedness remains unsettled. For models involving tidal interactions or minor mergers, there is an expected link between lopsidedness and the local environment. The evidence in this regard has been mixed (e.g., \\citealt{wp04}; \\citealt{bc+05}; \\citealt{aj+06,aj+07}; \\citet{dc+07}). %\\citet{dc+07} have undertaken %the most comprehensive investigation so far of this issue. Using the %rotational asymmetry measure $A_{180}$ to study a sample of over 3000 %galaxies, they find that close pairs of galaxies are more asymmetric %than other galaxies and that the asymmetry increases as the pair %separation decreases. They conclude that these global asymmetries %trace recent tidal interactions or mergers. The investigations summarized above have all involved relatively small samples of galaxies, making it difficult to assess the overall distribution of asymmetry or lopsidedness as a function of the basic parameters that characterize the structure of galaxies. This is the first of three papers in which we use the wealth of data available from the Sloan Digital Sky Survey (SDSS) to extend these studies of small samples (of-order one hundred galaxies) to large samples (tens of thousands). In this paper, we describe our sample selection and methodology. We also relate lopsidedness to the basic structural properties of the galaxies. In Paper II we will investigate the connection between lopsidedness and both star formation and black hole growth in galaxies. Finally, in Paper III we will examine the connection between lopsidedness and the local galaxy environment. In \\S\\ref{sec:data}, we begin by presenting an initial low-redshift sample from the SDSS and describe the observations and properties for its galaxies. Next, we explain our lopsidedness calculation. In \\S\\ref{sec:syserr}, we address the major data quality issues that limit the reliability of the measurements for portions of the sample. On this basis, we apply cuts on the observational parameters to weed out the problematic cases for our subsequent analysis. Next, \\S\\ref{sec:lopprops} describes the lopsidedness of galactic light distributions in different optical/near-IR bands, its correspondence with lopsided mass distributions, and its radial dependence. We then examine the relationship between lopsidedness and the basic structural properties of galaxies in \\S\\ref{sec:lopsfh}. Finally, we summarize our findings in \\S\\ref{sec:summary}. ", "conclusions": "} We have measured large-scale galactic asymmetry for a large sample of low-redshift ($z <$ 0.06) galaxies drawn from the Sloan Digital Sky Survey. Our use of lopsidedness, a radially averaged $m=1$ azimuthal Fourier mode, has proven useful for a large fraction of the sample. Images of a minority of galaxies in the sample have poor observational properties that cause significant systematic errors in the lopsidedness calculation, and these galaxies were removed from the sample via cuts on angular size, signal-to-noise, and ellipticity/inclination. Those cuts removed a higher fraction of the high-mass, high-mass-density, and high-concentration galaxies than those with low values of these structural properties. Nonetheless, the resulting sample is well represented by the galaxies of the same range of structural properties as the original sample. We find that there are no systematic differences between the $(g-i)$ colors of the brighter and fainter sides of lopsided galaxies. This implies that there is no systematic difference in the mass/light ratio \\citep{k+07}, and hence that the lopsided light distributions are primarily caused by lopsided distributions in the stellar mass. We have verified this through analysis of the relationship between color and mass/light ratio for both model galaxy spectral energy distributions and SDSS galaxy data. However, for our sample the lopsidedness in the $g$-band tends to be slightly greater than in the $r$- and $i$-bands. Thus, some of the lopsidedness in the light does arise from the effects of star-formation and/or dust extinction (which will more strongly affect the $g-$band light). Lopsidedness is a structural property that depends strongly on other structural properties. Galaxies with progressively lower concentration, stellar mass, or stellar mass density tend to have progressively higher lopsidedness. We show that the strongest and most fundamental correlation is between lopsidedness and stellar mass density. We also find that lopsidedness increases systematically with increasing radius, particularly for late-type galaxies. Lopsidedness can be induced through tidal stress associated with interactions with a companion galaxy or through accretion or minor mergers (e.g. \\citealt{zr97}; \\citealt{bc+05}). Galaxies with low density will be most affected by tidal stress, and the effects of a tidal perturbation will last longer in such systems due to the longer dynamical times. The same arguments pertain to the outer parts of galaxies. Thus, the two above results make good physical sense. Alternatively, if the dark matter halo is lopsided, its effects on the structure of the stellar disk will be more pronounced in the outer region and in galaxies with low mass and low density (where dark matter is more dynamically important). The relatively large values of lopsidedness we measure to be commonplace ($A_1 > 0.1$) appear to be too large to be generated by internally generated dynamical processes (e.g., \\citealt{mt97}). Our overall goal in this investigation has been to use lopsidedness as a way of quantifying the signature of moderate or weak global dynamical perturbations. The next step will be to determine the connections between such perturbations and both the on-going/recent star formation and the growth of supermassive black holes in galaxies. These connections can help constrain the processes and conditions that guide the formation and evolution of the galaxies. In future papers we will address these questions using the present sample of galaxies. %are also thought to induce bursts of star formation. Using two well %studied SFH indicators, we have shown that lopsidedness is typically %associated with bursty and fast star formation, and quiescent star %formation is typical in symmetric galaxies. We will extend this %discussion in future work. %Interactions are also thought to induce increased star formation rates %in galaxies. We have found that lopsided galaxies produce stars at %much faster rates per unit stellar mass than symmetric galaxies. %Lopsidedness is then another way to qualitatively describe the stellar %population age of galaxies. Taking lopsidedness as a signature of a %recent interaction, the link between star formation and lopsidedness %may be that they are symptoms of the same interactions. We will %extend this discussion in future work. JB acknowledges the receipt of a FCT post-doctoral grant BPD/14398/2003. We would like to thank Vivienne Wild for reading a draft of the manuscript. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington." }, "0710/0710.4547_arXiv.txt": { "abstract": "We present an extensive data set of $\\sim 150$ localized features from \\Cassit{} images of Saturn's Ring~A, a third of which are demonstrated to be persistent by their appearance in multiple images, and half of which are resolved well enough to reveal a characteristic ``propeller'' shape. We interpret these features as the signatures of small moonlets embedded within the ring, with diameters between 40 and 500~meters. The lack of significant brightening at high phase angle indicates that they are likely composed primarily of macroscopic particles, rather than dust. With the exception of two features found exterior to the Encke Gap, these objects are concentrated entirely within three narrow ($\\sim 1000$ km) bands in the mid-A Ring that happen to be free from local disturbances from strong density waves. However, other nearby regions are similarly free of major disturbances but contain no propellers. It is unclear whether these bands are due to specific events in which a parent body or bodies broke up into the current moonlets, or whether a larger initial moonlet population has been sculpted into bands by other ring processes. ", "introduction": "Saturn's main rings (particularly Ring~A) were determined by the Voyager Radio Science experiment \\citep{Zebker85} to be primarily composed of a distribution of icy particles of diameter $D \\gtrsim 1$~cm. A steep cutoff in the size-distribution was discerned at $D \\sim 20$~m, but particles larger than that value could not be probed due to limitations imposed by the radio experiment's carrier wavelength. At the large end of the particle-size distribution are the two known moonlets embedded in gaps in the outer A~Ring, Pan and Daphnis, of diameters $\\sim 28$~km and $\\sim 8$~km, respectively \\citep{PorcoSci07}. Nothing was known about the distribution of intermediate-size ring particles, between 20~m and 8~km in diameter, until the first evidence of ``missing-link'' particles was found in very-high-resolution images taken by the \\Cassit{} spacecraft during its insertion into Saturn orbit \\citep{Propellers06}. These intermediate-size moonlets are not directly seen, but rather the propeller-shaped disturbances they create in the ring continuum. This morphology had previously been predicted by numerical simulations \\citep{SS00,SSD02,Seiss05}. The sizes of the perturbing moonlets, subject to some ambiguities in interpretation, were given as $D \\sim 100$ meters. Their surface densities are quite low relative to smaller particles, corroborating the steep cutoff reported by \\citet{Zebker85}. We here present and analyze a data set of 158 localized features in the A~Ring, many of which are well-enough resolved to reveal the characteristic propeller shape. Four of these objects are those reported by \\citet{Propellers06}, and another eight were first noted by \\citet{Sremcevic07}. Recently, \\citet{Espo07} have found evidence for similarly-sized moonlets in the narrow and highly disturbed F~Ring; however, this is not directly applicable to our analysis because there is no particular reason to expect the particle-size distribution of the F~Ring to be simply related to that of the A~Ring. Section~\\ref{Propellers} gives further background on the nature and interpretation of propellers. Section~\\ref{Observations} summarizes the imaging sequences we used in compiling our data set, and Section~\\ref{Analysis} describes the process by which features were identified and characterized with one of two models (``resolved'' or ``unresolved''). Our results are summarized in Section~\\ref{Results}. Section~\\ref{Interpretation} contains further discussion of the interpretation of propeller features. A full and unabridged presentation of our data set is given as an Appendix. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=10cm,keepaspectratio=true]{f1.pdf} \\caption{Particle trajectories under Hill's equations \\citep[see, e.g.,][ch.~3.13]{MD99}, showing the propeller-shaped chaotic zones as well as associated moonlet wakes. Radial ($r$)) and azimuthal ($\\ell$) coordinates are each shown in units of Hill radii; the perturbing mass is located at the origin, and the $L_1$ and $L_2$ Lagrange points are at $\\ell = 0$, $r = \\pm 1$ The direction towards Saturn is down, and the orbital direction is to the right. Scattered trajectories deflected by $>5R_H$ are not shown. This simple model, which neglects inter-particle collisions as well as self-gravity, is presented for conceptual purposes only. \\label{Illustration}} \\end{center} \\end{figure} ", "conclusions": "" }, "0710/0710.1648_arXiv.txt": { "abstract": "We have carried out a search for radio emission from six X-ray dim isolated neutron stars (XDINSs) observed with the Robert C. Byrd Green Bank Radio Telescope (GBT) at 820~MHz. No bursty or pulsed radio emission was found down to a $4\\sigma$ significance level. The corresponding flux limit is 0.01--0.04~mJy depending on the integration time for the particular source and pulse duty cycle of 2\\%. These are the most sensitive limits yet on radio emission from these objects. ", "introduction": "The \\emph{ROSAT} mission discovered a group of seven nearby low-luminosity isolated neutron stars, termed the X-ray dim isolated neutron stars (XDINSs). They share very similar properties and are characterized by soft blackbody-like spectra in the range $\\sim 40$--100 eV, very faint optical counterparts ($V>25$), long spin periods of 3--12~s (see Table~\\ref{table}). For a recent review see \\citet{haberl2007}. So far, no confident detections of pulsed radio emission were found from XDINSs. The aim of this project is to search for pulsed and bursty radio emission from XDINSs to link them in their evolutionary scenarios with other classes of neutron stars, such as magnetars and rotating radio transients (RRATs). All three populations of neutron stars have similar properties, such as period, period derivative, age, and magnetic field so that connections between them are plausible. ", "conclusions": "The sporadicity of the RRATs' radio emission led to immediate suggestions that they are related to other classes of traditionally ``radio-quiet'' neutron stars such as XDINSs and magnetars. \\citet{popov2006} have shown that the implied birthrate of RRATs is more consistent with that of XDINSs than that of magnetars. As shown in the P-\\.P diagram on the Figure~\\ref{ppdot}, RRATs and XDINSs also have similar periods and period derivatives, implied ages and magnetic fields. However, the RRATs spin-down properties are also consistent with those of the normal pulsar population and X-ray observations of one RRAT~\\citep{mmclaugh2007} reveal properties similar to those of both normal radio pulsars and XDINSs. \\begin{figure} \\includegraphics[scale=0.44]{p-pdot_bw.ps} \\caption{P-\\.P diagram. Lines on the top mark the period values for the objects with unknown yet period derivatives: seven RRATs (solid), one magnetar SGR~1627$-$41 and two candidates AX~J1845$-$03 and CXO~J164710.2$-$455216 (dashed), and one XDINS RX~J1605.3+3249 (longer dashed). } \\label{ppdot} \\end{figure} The RRATs are powerful sources of isolated radio bursts, and we have not detected such bursts, or any periodic emission, from the six XDINSs we have observed. Because the distances to the XDINSs are believed to be much smaller than those to the RRATs, we should have had high sensitivity to RRAT-like radio emission. However, our non-detection of such emission does not necessarily mean that there is no relationship between these two source classes. XDINSs may simply be ``radio-quiet'', but it is also likely that perhaps the narrow radio beams from these XDINSs are simply misaligned with our line-of-sight. It is possible that searches at lower frequencies, where radio emission beams are believed to be wider, may be more sensitive to radio emission from XDINSs. Indeed, Malofeev and co-authors~\\citep{malofeev2005, malofeev2007} reported detection of radio emission from RX~J1308.6+2127 and RX~J2143.7+0654 at the low frequency of 111~MHz. On the other hand, XDINSs could have very steep spectral indices. If the detection of Malofeev and co-authors is real, our non-detection of radio emission from these two XDINSs at 820~MHz sets a lower limit on the spectral index of 3.6. Finally, it is possible that our non-detection of radio emission from these XDINSs is due to the large amount of contamination from RFI. Our search highlights the importance of improved excision algorithms for impulsive, broadband terrestrial interference. \\begin{theacknowledgments} SZ thanks STFC (ex-PPARC) for support through an AF. The Robert C. Byrd Green Bank Telescope (GBT) is operated by the National Radio Astronomy Observatory which is a facility of the U.S. National Science Foundation operated under cooperative agreement by Associated Universities, Inc. \\end{theacknowledgments}" }, "0710/0710.1683.txt": { "abstract": "The lensing cross section of triaxial halos depends on the relative orientation between a halo's principal axes and its line of sight. Consequently, a lensing subsample of randomly oriented halos is not, in general, randomly oriented. Using an isothermal mass model for the lensing galaxies and their host halos, we show that the lensing subsample of halos that produces doubles is preferentially aligned along the lines of sight, whereas halos that produce quads tend to be projected along their middle axes. These preferred orientations result in different projected ellipticity distributions for quad, doubles, and random galaxies. We show that $\\approx 300$ lens systems must be discovered to detect this effect at the $95\\%$ confidence level. We also investigate the importance of halo shape for predicting the quad-to-double ratio and find that the latter depends quite sensitively on the distribution of the short-to-long axis ratio, but is otherwise nearly independent of halo shape. Finally, we estimate the impact of the preferred orientation of lensing galaxies on their projected substructure mass fraction, and find that the observed alignment between the substructure distribution and the mass distribution of halos result in a negligible bias. ", "introduction": "Statistics of lensing galaxies have been used as cosmological and galaxy formation probes since early in the modern history of gravitational lensing \\citep[][]{turneretal84}. Lensing rates can be used to constrain dark energy \\citep[][]{fukugitaetal92, chae03,mitchelletal05, chae07,ogurietal07}, to probe the structure of lensing galaxies \\citep[][]{keeton01d,kochanekwhite01, chae05}, and to probe galaxy evolution \\citep[][]{chaemao03, ofeketal03,rusinkochanek05}. While the use of lensing statistics as a cosmological probe has had mixed success, particularly early on, it remains a unique probe with entirely different systematics from more traditional approaches. Consequently, lensing statistics are likely to remain a fundamental cross-check of our understanding of cosmology and galaxy evolution. One of the difficulties that confronts the study of lensing statistics is that, in general, the halo population that produces gravitational lenses can in fact be a highly biased subsample of the general halo population. For instance, it has long been known that while early type galaxies compose only $\\approx 30\\%$ of all luminous galaxies, the majority of lensing galaxies are in fact early type since these tend to be more massive and reside in more massive halos than their late counterparts. By the same token, lensing early type galaxies tend to have higher luminosity and velocity dispersions than non-lensing early type galaxies \\citep[][]{moelleretal06, boltonetal06}. Overall, then, when interpreting lensing statistics, one ought to always remember that by selecting lensing galaxies one is automatically introducing an important selection effect that can significantly bias the distribution of any galaxy observable that has an impact on the lensing probabilities. Here, we consider one such source of bias, the triaxiality of galaxy halos.\\footnote{Throughout this work, we will be using the term galaxy and halo more or less interchangeably. The reason for this is that we are primarily focused on the impact of halo triaxiality on the lensing cross section, and the latter depends only on the {\\it total} matter density. Consequently, differentiating between halo and galaxy would only obfuscate presentation and introduce unnecessary difficulties. For instance, while modeling the total matter distribution as isothermal is a reasonable approximation, neither the baryons nor the dark matter by itself is isothermally distributed. Thus, it is much simpler to adopt an isothermal model, and refer to the baryons plus dark matter as a single entity, than to try to differentiate between the two. Likewise, when discussing triaxiality, what is important in this work is the triaxiality of the total matter distribution.} That halo triaxiality can have important consequences for lensing statistics has been known for several years. For instance, \\citet[][]{ogurikeeton04} have shown that triaxiality can significantly enhance the optical depth of large image separation lenses. Similar conclusions have been reached concerning the formation of giant arcs by lensing clusters \\citep[see e.g.][and references therein]{ogurietal03, rozoetal06c, hennawietal07}. Curiously, however, little effort has gone into investigating how observational properties of lensing galaxies can be different from those of the galaxy population as a whole due to the triaxial structure of galactic halos. This work addresses this omission. The first observable we consider is the projected axis ratio of lensing galaxies. Roughly speaking, given that non-zero ellipticities are needed in order to produce quad systems, one would generically expect lenses that lead to this image configuration to be more elliptical than the overall galaxy population. Likewise, lensing galaxies that produce doubles should, on average, be slightly more circular than a random galaxy. There can, however, be complications for these simple predictions due to halo triaxiality. For instance, given a prolate halo, projections along the long axis of the lens will result in highly concentrated, very circular profiles. Will the increase in Einstein radius of such projections compensate for the lower ellipticity of the system, implying most quads will be projected along their long axis, or will it be the other way around? Clearly, the relation between ellipticity and lensing cross sections is not straightforward once triaxiality of the lensing galaxies is taken into account, but it seems clear that there should be some observable difference between the ellipticity distribution of lensing galaxies and that of all early types. Interestingly, no such difference has been observed \\citep[][]{keetonetal97,rusintegmark01}, which seems to fly in the face of our expectations \\citep[though see also the discussion in][]{keetonetal98}. Is this actually a problem, or will a quantitative analysis show that the consistency of the two distributions is to be expected? Here, we explicitly resolve this question, and demonstrate that current lens samples are much too small to detect the expected differences. Having considered the ellipticity distribution of random and lensing galaxies, it is then a natural step to investigate the impact of halo triaxiality on predictions of the quad-to-double ratio. Specifically, it is well known that the quad-to-double ratio is sensitive to the ellipticity distribution of lensing galaxies \\citep[][]{keetonetal97}, so if lensing can bias the distribution of ellipticities in lensing galaxies, then it should also affect the predicted quad-to-double ratios. This is an important point because it has been argued that current predictions for the quad-to-double ratio are at odds with observations. More specifically, the predicted quad-to-double ratio for the CLASS \\citep[Cosmic Lens All-Sky Survey,][]{myersetal03,browneetal03} sample of gravitational lenses is too low relative to observations \\citep[][]{rusintegmark01,hutereretal05}. Curiously, however, recent work on the quad-to-double ratio observed in the SQLS \\citep[Sloan Digital Sky Survey Quasar Lens Search,][]{ogurietal06,inadaetal07}. suggests that the exact opposite is true for the latter sample, namely, theoretical expectations are too high relative to observations \\citep[][]{oguri07}. In either case, it is of interest to determine how exactly does triaxiality affects theoretical predictions, especially since the aforementioned difficulties with the CLASS sample has led various authors to offer possibilities as to how one might boost the expected quad-to-double ratios. Specifically, one can boost the quad-to-double ration in the class sample either from the effect of massive satellite galaxies near the lensing galaxies \\citep[][]{cohnkochanek04}, or through the large-scale environment of the lensing galaxy \\citep[][]{keetonzabludoff04}. Clearly, we should determine whether halo triaxiality can be added to this list. This brings us then to the final problem we consider here, namely whether the substructure population of lensing galaxies is different from that of non-lensing galaxies. Specifically, we have argued that lensing galaxies will not be isotropically distributed in space. Since the substructure distribution of a dark matter halo is typically aligned with its parent halo's long axis \\citep[][]{zentneretal05,libeskindetal05,agustssonbrainerd06,azzaroetal06}, it follows that the projected distribution of substructures for lensing galaxies may in fact be different for lensing halos than for non-lensing halos. Such an effect could be quite important given the claimed tension between the Cold Dark Matter (CDM) predictions for the substructure mass fraction of halos \\citep[see][]{maoetal04} and their observed values \\citep[][]{dalalkochanek02a,kochanekdalal04}. Likewise, such a bias would impact the predictions for the level of astrometric and flux perturbations produced by dark matter substructures in gravitational lenses \\citep[][]{rozoetal06,chenetal07}. Here, we wish to estimate the level at which the projected substructure mass fraction of lensing halos could be affected due to lensing biasing. The paper is organized as follows: in section \\ref{sec:biases} we derive the basic equations needed to compute how observable quantities will be biased in lensing galaxy samples due to halo triaxiality. Section \\ref{sec:model} presents the model used in this work to quantitatively estimate the level of these biases, and discusses how lensing halos are oriented relative to the line of sight as a function of the halos' axes ratios. Section \\ref{sec:axis} investigates the projected axis ratio distributions of lensing versus non-lensing galaxies, and demonstrates that present day lensing samples are too small to detect the triaxiality induced biases we have predicted. Section \\ref{sec:ratio} discusses the problem of the quad to double ratio, and section \\ref{sec:subs} demonstrates that halo triaxiality biases the projected substructure mass fraction in lensing halos by a negligible amount. Section \\ref{sec:caveats} discusses a few of the effects we have ignored in our work and how these may alter our results, and finally section \\ref{sec:summary} summarizes our work and presents our conclusions. %---------------------------------------------- ", "conclusions": "\\label{sec:summary} The triaxial distribution of mass in galactic halos implies that the probability that a galaxy becomes a lens is dependent on the relative orientation of the galaxy's major axis to the line of sight. Consequently, a subsample of randomly oriented galaxies that act as strong lenses will {\\it not} be randomly oriented in space. The relative orientation and the strength of the alignment depends on the shape of the matter distribution, and on the type of lens under consideration: prolate doubles have a high probability of being project along their long axis, whereas the distribution of oblate doubles is nearly isotropic. Prolate quads are most often projected along their middle axis, though the degree to which alignment occurs is not as strong as for prolate doubles. Interestingly, highly prolate quads are also more likely to be projected along their short axis than along the long axis, though this very quickly changes as halos become more triaxial and less prolate. Oblate quads strongly avoid projections along the short axis of the lens, but projections along the other two axis are almost equally likely. An important consequence of the differences in the distribution of halo orientations for quad lenses, double lenses, and the galaxy population as a whole is that the ellipticity distribution of these various samples must be different, even if the distribution of halo shapes is the same. Specifically, we predict that quad lenses are typically more elliptical than random galaxies, and that the ellipticity distribution of doubles is very slightly more circular than that of random galaxies. While current data do not show any indication of these trends, we have shown that $\\approx 300\\ (1,400)$ lenses are necessary to obtain a $2\\sigma\\ (5\\sigma)$ detection of the effect. The fact that halo triaxiality affects the ellipticity distribution of lensing galaxies also means that halo triaxiality needs to be properly taken into account in lensing statistics. Consequently, we estimate how the biased lensing cross sections of galaxies depend on halo shape, and find that they are nearly independent of the halo shape parameter $T$. Instead, the mean biased cross section of a lens depends almost exclusive on the distribution on the short-to-long axis ratio $q_2$ (often denoted by $s$). Finally, given that the distribution of substructures in numerical simulations is observed to be preferentially aligned with the long axis of the host halos, we estimate how the preferred orientation of lensing galaxies affects their predicted substructure mass fraction. We find that biases due to non-isotropic distribution of halos relative to the line of sight have an insignificant impact on the mean substructure mass fraction of lensing galaxies. {\\bf Acknowledgements: } ER would like to thank Christopher Kochanek for numerous discussions and valuable comments on the manuscript which have greatly improved both the form and content of this work. The authors would also like to thank to Emilio Falco for kindly providing the isophotal axis ratio data that was needed for producing Figure \\ref{fig:cumq}, and to Charles Keeton for a careful reading of the manuscript. ER was funded by the Center for Cosmology and Astro-Particle Physics (CCAPP) at The Ohio State University. ARZ has been funded by the University of Pittsburgh, the National Science Foundation (NSF) Astronomy and Astrophysics Postdoctoral Fellowship program through grant AST 0602122, and by the Kavli Institute for Cosmological Physics at The University of Chicago. This work made use of the National Aeronautics and Space Administration Astrophysics Data System. % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- %--------------------------------------------------------------------------------- % -------------------------------------------------------------------------------- % --------------------------------------------------------------------------------" }, "0710/0710.5407_arXiv.txt": { "abstract": "{Based on the analogy with non-minimal $SU(2)$ symmetric Wu-Yang mono\\-pole with regular metric, the solution describing a non-minimal $U(1)$ symmetric Dirac monopole is obtained. In order to take into account the curvature coupling of gravitational and electromagnetic fields, we reconstruct the effective metrics of two types: the so-called associated and optical metrics. The optical metrics display explicitly that the effect of birefringence induced by curvature takes place in the vicinity of the non-minimal Dirac monopole; these optical metrics are studied analytically and numerically. } \\email 1 {Alexander.Balakin@ksu.ru}% \\email 2 {Alexei.Zayats@ksu.ru}% ", "introduction": "The Dirac monopole as a specific static spherically symmetric solution to the minimal Einstein-Maxwell equations has became a subject of discussion in tens papers, reviews and books (see, e.g., \\cite{Dirac,000,00,0,1,3,2}). The motion of massive and massless particles, which possess electric charge or are uncharged, is studied in detail (see, e.g., \\cite{4,5} and references therein). In the paper \\cite{BaZa07} we introduced and discussed the $SU(2)$ symmetric Wu-Yang monopole of the new type, namely, the non-minimal monopole with regular metric. Since the non-minimal Wu-Yang monopole is effectively Abe\\-lian, it is naturally to consider the corresponding analog of that solution in the framework of non-minimal electro\\-dynamics. Mention that non-minimal models with mag\\-netic charge have been discussed earlier (see, e.g., \\cite{Horn,MHS}), but the exact analytical regular solution obtained here as direct reduction to the $U(1)$ symmetry is the new one. The second novelty of the presented paper is the investigation of photon dynamics in the vicinity of the Dirac monopole accounting the curvature coupling of the gravitational and electromagnetic fields. In the presence of non-minimal interaction (induced by curva\\-ture) the master equations for electromagnetic and gravitational fields in vacuum can be rewritten as the master equations in some effective anisotropic (quasi)me\\-dium \\cite{B1,BL05}. This means that two effective (optical) metrics can be introduced \\cite{HehlObukhov,Perlick,BZ05}, so that the photon propagation in vacuum interacting with curvature is equivalent to the photon motion in the effective space-time with the first or second optical metric, depending on the photon polarization. Even if the real space-time has a regular metric, the optical metrics can be singular, admitting the interpretation in terms of the so-called ``trapped surfaces'' and ``inaccessible zones'' \\cite{AG1,AG2}. We discuss this problem in Sect. 3. Numerical modeling of the photon orbits, presented in Sect. 4, supplement our conclusions. ", "conclusions": "Qualitative and numerical analysis of the photon orbits in the vicinity of non-minimal Dirac monopole with regular metric has demonstrated the following interesting features. 1. Propagation of the electromagnetic waves in the vicinity of non-minimal Dirac monopole is characterized by birefringence, induced by curvature, i.e., the phase velocities of waves depend on their polarization. Two different optical metrics should be introduced to describe two principal states of polarization. 2. The metric of the non-minimal Dirac monopole, obtained and discussed in this paper, is regular, thus, all the singularities of the optical metrics have a dynamic origin and are supported by the non-minimal (curvature induced) interaction of the gravitational and electromagnetic fields. 3. The points of self-intersection, the points of the closest approach, the reverse points, etc. in the photon trajectories can be recognized and catalogued for different combinations of the values of the impact parameter and the parameter of non-minimal coupling. We discussed here only the principal pictures. \\Acknow This work was partially supported by the DFG through project No. 436RUS113/487/0-5. The authors are grateful to Claus L\\\"ammerzahl for valuable remarks. \\small" }, "0710/0710.5294_arXiv.txt": { "abstract": "{Thanks to remarkable progress, radial velocity surveys are now able to detect terrestrial planets at habitable distance from low-mass stars. Recently, two planets with minimum masses below 10~$M_{\\oplus}$have been reported in a triple system around the M-type star Gliese 581. These planets are found at orbital distances comparable to the location of the boundaries of the habitable zone of their star. } {In this study, we assess the habitability of planets Gl~581c and Gl 581d (assuming that their actual masses are close to their minimum masses) by estimating the locations of the habitable-zone boundaries of the star and discussing the uncertainties affecting their determination. An additional purpose of this paper is to provide simplified formulae for estimating the edges of the habitable zone. These may be used to evaluate the astrobiological potential of terrestrial exoplanets that will hopefully be discovered in the near future.} {Using results from radiative-convective atmospheric models and constraints from the evolution of Venus and Mars, we derive theoretical and empirical habitable distances for stars of F, G, K, and M spectral types. } {Planets Gl~581c and Gl~581d are near to, but outside, what can be considered as the conservative habitable zone. Planet `c' receives 30\\% more energy from its star than Venus from the Sun, with an increased radiative forcing caused by the spectral energy distribution of Gl~581. This planet is thus unlikely to host liquid water, although its habitability cannot be positively ruled out by theoretical models due to uncertainties affecting cloud properties and cloud cover. Highly reflective clouds covering at least 75\\% of the day side of the planet could indeed prevent the water reservoir from being entirely vaporized. Irradiation conditions of planet `d' are comparable to those of early Mars, which is known to have hosted surface liquid water. Thanks to the greenhouse effect of CO$_2$-ice clouds, also invoked to explain the early Martian climate, planet `d' might be a better candidate for the first exoplanet known to be potentially habitable. A mixture of several greenhouse gases could also maintain habitable conditions on this planet, although the geochemical processes that could stabilize such a \\textit{super-greenhouse} atmosphere are still unknown. } {} ", "introduction": "The M-type star Gl~581 hosts at least 3 planets, which were detected using radial velocity measurements by Bonfils et al. \\citeyearpar{2005A&A...443L..15B} (planet 'b') and Udry et al. \\citeyearpar{2007A&A...469L..43U} (planets `c' and `d'). The properties of this star and its planets are given in Table~\\ref{tab:gl581}. Before this discovery, only two exoplanets were known to have a minimum mass below 10~$M_{\\oplus}$, which is usually considered as a boundary between terrestrial and giant planets, the latter having a significant fraction of their mass in an H$_2$-He envelope. The first one was GJ~876d, a very hot planet ($P\\leq 2$~days) with a minimum mass of 5.9~M$_{\\oplus}$ \\citep{2005ApJ...634..625R}. The other one is OGLE-05-390L b, found to be a $\\sim$5.5~M$_{\\oplus}$ cold planet at 2.1~AU from its low-mass parent star thanks to a microlensing event \\citep{2006Natur.439..437B,2006ApJ...651..535E}. Neither of these two planets is considered as habitable, even with very loose habitability criteria. In the case of Gl~581, and as already mentioned by Udry et al. (2007), the locations of planet `c' and `d' must be fairly close to the inner and outer edges, respectively, of the habitable zone (HZ). In this paper, we investigate the atmospheric properties that would be required to make the habitability of these planets possible. Because of its equilibrium temperature of $\\sim$300~K when calculated with an albedo of 0.5, it has been claimed that the second planet of this system, Gl~581c, is potentially habitable (Udry et al. 2007), with climatic conditions possibly similar to those prevailing on Earth. After a brief discussion about the relationship between the equilibrium temperature and habitability, we summarize in this paper what are usually considered as the boundaries of the circumstellar HZ and the uncertainties on their precise location. In Sect.~\\ref{sec:HZstars} we provide parameterizations to determine such limits as a function of the stellar luminosity and effective temperature. These can be used to evaluate the potential habitability of the terrestrial exoplanets that should soon be discovered. We then discuss the specific case of the system around Gl~581. \\begin{table}[!t] \\caption{Properties of the star Gl~581 and its 3 detected planets, from Udry et al. (2007).} \\begin{center} \\begin{tabular}{lllll} \\hline Star & $T_{\\rm eff}$ (K) &$M$/M$_{\\odot}$ & $R$/R$_{\\odot}$ & $L$/L$_{\\odot}$ \\\\ \\hline Gl~581 & 3200 &0.31 & 0.38 & 0.0135 \\\\ \\hline &&& \\\\ &&& \\\\ \\hline Planets & $a$~(AU) & $M_{\\rm min}$/M$_{\\oplus}$ & $R_{\\rm min}$/R$_{\\oplus}$ & stellar flux\\\\ & &$^{*}$ & $^{**}$ & $S/S_{0}$$^{***}$\\\\ \\hline b & 0.041 & 15.6 & 2.2-2.6 & 8.1 \\\\ \\rowcolor{lightgray} c & 0.073 & 5.06 & 1.6-2.0 & 2.55 \\\\ \\rowcolor{lightgray} d & 0.253 & 8.3 & 1.8-2.2 & 0.21 \\\\ \\hline \\end{tabular} \\end{center} The potential habitability of planets `c' and `d', highlighted in grey, is discussed in this paper. \\\\ $^{*}$ $M_{\\rm min}=M \\sin i$, where $i$ is the orbital inclination.\\\\ $^{**}$ Radius for a rocky and ocean planet, respectively \\citep{2007Sotin,2007ApJ...665.1413V}.\\\\ $^{***}$ $S_0$ is the solar flux at 1 AU: 1360 W m$^{-2}$.\\\\ \\label{tab:gl581} \\end{table} ", "conclusions": "According to our present knowledge, based on available models of planetary atmospheres, and assuming that the actual masses of the planets are the minimum masses inferred from radial velocity measurements, Gl~581c is very unlikely to be habitable, while Gl~581d could potentially host surface liquid water, just as early Mars did. Because of the uncertainties in the precise location of the HZ boundaries, planets at the edge of what is thought to be the HZ are crucial targets for future observatories able to characterize their atmosphere. At the moment, our theory of habitability is only confirmed by the divergent fates of Venus and the Earth. We will have to confront our models with actual observations to better understand what makes a planet habitable. The current diversity of exoplanets (planets around pulsars, hot Jupiters, hot Neptunes, super-Earths, etc) has already taught us that Nature has a lot more imagination when building a variety of worlds than we expected from our former models inspired by the Solar System. It is obvious that the idealized model of a habitable planet atmosphere, where the two important constituents are CO$_2$ and H$_2$O, CO$_2$ being controlled by the carbonate-silicate cycle, is likely to represent only a fraction of the diversity of terrestrial planets that exist at habitable distances from their parent star. As an example, planets fully covered by an ocean may be common, either because they are richer in water than Earth or because the distribution between surface and mantle water is different, or perhaps simply because, for a given composition, the mass-to-surface ratio and thus the water-to-surface ratio increases with the planetary mass, as noted by Lissauer \\citeyearpar{1999Lissauer}. Without emerged continents, it is not at all clear that the carbonate-silicate cycle could operate. The planets around Gl~581 can fall into this category since they are significantly more massive than the Earth (especially the $>$8~$M_{\\oplus}$ planet Gl~581d) and also because they may have started their formation in the outer and more water-rich region of the protoplanetary disk. Darwin/TPF-I and TPF-C could eventually reveal what the actual properties of the atmosphere of Gl~581c and Gl~581d are. From their thermal light curves we could infer if a thick atmosphere is making the climate more or less uniform on both the day and night hemispheres of these planets, despite a (nearly?) synchronized rotation \\citep{2004ASPC..321..170S}. Visible and mid-IR water vapor bands could be searched in the atmosphere of Gl~581d to confirm its habitability. Mid-IR spectra of this planet could also reveal other greenhouse gases at work. Spectral observations of Gl~581c could potentially distinguish between a Venus-like atmosphere dominated by CO$_2$ or an H$_2$O-rich atmosphere. The detection of O$_2$ on this planet would generate a fascinating debate about its possible origin: as either a leftover of H$_2$O photolysis and H escape or a biological release. There is certainly no doubt that Gl~581c and Gl~581d are prime targets for exoplanet characterization missions." }, "0710/0710.0364_arXiv.txt": { "abstract": "We explore the cosmological consequences of Modified Gravity (MOG), and find that it provides, using a minimal number of parameters, good fits to data, including CMB temperature anisotropy, galaxy power spectrum, and supernova luminosity-distance observations without exotic dark matter. MOG predicts a bouncing cosmology with a vacuum energy term that yields accelerating expansion and an age of $\\sim$13 billion years. ", "introduction": "The preferred model of cosmology today, the $\\Lambda$CDM model, provides an excellent fit to cosmological observations, but at a substantial cost: according to this model, {\\em about 95\\% of the universe is either invisible or undetectable, or possibly both} \\cite{Komatsu2008}. This fact provides a strong incentive to seek alternative explanations that can account for cosmological observations without resorting to dark matter or Einstein's cosmological constant. For gravitational theories designed to challenge the $\\Lambda$CDM model, the bar is set increasingly higher by recent discoveries. Not only do such theories have to explain successfully the velocity dispersions, rotational curves, and gravitational lensing of galaxies and galaxy clusters, the theories must also be in accord with cosmological observations, notably the acoustic power spectrum of the cosmic microwave background (CMB), the matter power spectrum of galaxies, and the recent observation of the luminosity-distance relationship of high-$z$ supernovae, which is seen as evidence for ``dark energy''. Modified Gravity (MOG) \\citep{Moffat2006a} has been used successfully to account for galaxy cluster masses \\citep{Brownstein2006b}, the rotation curves of galaxies \\citep{Brownstein2006a}, velocity dispersions of satellite galaxies \\citep{Moffat2007}, and globular clusters \\citep{Moffat2007a}. It was also used to offer an explanation for the Bullet Cluster \\citep{Brownstein2007} without resorting to cold dark matter. Remarkably, MOG also meets the challenge posed by cosmological observations. In this paper, it is demonstrated that MOG produces an acoustic power spectrum, a matter power spectrum, and a luminosity-distance relationship that are in good agreement with observations, and require no dark matter nor Einstein's cosmological constant. In the arguments presented here, we rely on simplified analytical calculations. We are not advocating these as substitutes for an accurate numerical analysis. However, a thorough numerical analysis requires significant time and resources; before these are committed, it is useful to be able to demonstrate if a theory is viable, and if the additional effort is warranted. In the next section, we review the key features of MOG. This is followed by sections presenting detailed calculations for the luminosity-distance relationship of high-$z$ supernovae, the acoustic power spectrum of the CMB, and the galaxy power spectrum. A concluding section summarizes our results and maps out future steps. ", "conclusions": "In this paper, we demonstrated how Modified Gravity can account for key cosmological observations using a minimum number of free parameters. Although MOG permits the running of its coupling and scaling constants with time and space, we made very little use of this fact. Throughout these calculations, we used consistently the value of $\\alpha\\simeq 19$ for the MOG coupling constant, consistent with a flat universe with $\\Omega_b\\simeq 0.05$ visible matter content, no dark matter, nor Einstein's cosmological constant. In nearly all cosmological calculations, we set the MOG scaling constant $\\mu$ to the inverse of the radius of the visible universe, which is a natural choice. The only exception is the mass power spectrum calculation, where the scaling constant enters in conjunction with the wave number $k$, and describes gravitational interactions between nearby concentrations of matter, not on the cosmological scale. The theory requires {\\em no other parameters} to obtain the remarkable fits to data that have been demonstrated here. At all times, $\\lim\\limits_{r\\rightarrow 0}G=G_N$, i.e., the effective gravitational constant at short distances remains Newton's constant of gravitation. For this reason, the predictions of MOG are {\\em not contradicting our knowledge of the processes of the initial nucleosynthesis}, taking place at redshifts of $z\\simeq 10^{10}$, since the interactions take place over distance scales that are much shorter than the horizon scale. Our calculations relied on analytical approximations. This is dictated by necessity, not preference. We recognize that numerical methods, including high-accuracy solutions of coupled systems of differential equations, as in {\\tt CMBFAST} \\citep{Seljak1996}, or $N$-body simulations, can provide superior results, and may indeed help either to confirm or to falsify the results presented here. Nevertheless, our present work demonstrates that at the very least, MOG provides a worthy alternative to $\\Lambda$CDM cosmology." }, "0710/0710.2157_arXiv.txt": { "abstract": "We investigate the incidence of major mergers creating ${\\rm M}_{\\rm star}>10^{11} {\\rm M}_{\\sun}$ galaxies in the dense environments of present-day groups and clusters more massive than ${\\rm M}_{\\rm halo}=2.5\\times10^{13}{\\rm M}_{\\sun}$. We identify 38 pairs of massive galaxies with mutual tidal interaction signatures selected from $>5000$ galaxies with ${\\rm M}_{\\rm star}\\geq5\\times10^{10} {\\rm M}_{\\sun}$ that reside in a halo mass-limited sample of 845 groups. We fit the images of each galaxy pair as the line-of-sight projection of symmetric models and identify mergers by the presence of residual asymmetric structure associated with both progenitors, such as nonconcentric isophotes, broad and diffuse tidal tails, and dynamical friction wakes. At the resolution and sensitivity of the SDSS, such mergers are found in 16\\% of the high-mass, galaxy-galaxy pairs with $\\leq1.5$ $r$-band magnitude differences and $\\leq30$ kpc projected separations. Relying on automated searches of major pairs from the SDSS spectroscopic galaxy sample will result in missing 70\\% of these mergers owing to spectroscopic incompleteness in high-density regions. We find that 90\\% of these mergers are between two nearly equal-mass progenitors with red-sequence colors and centrally-concentrated morphologies, in agreement with numerical simulations that predict that an important mechanism for the formation of massive elliptical galaxies is the dissipationless (gas-poor or so-called dry) major merging of spheroid-dominated galaxies. We identify seven additional ${\\rm M}_{\\rm star}>10^{11} {\\rm M}_{\\sun}$ mergers with disturbed morphologies and semi-resolved double nuclei. Mergers at the centers of massive groups are more common than between two satellites, but both types are morphologically indistinguishable and we tentatively conclude that the latter are likely located at the dynamical centers of large subhalos that have recently been accreted by their host halo, rather than the centers of distinct halos seen in projection. We find that the frequency of central and satellite merging diminishes with group mass in a manner that is consistent with dynamical friction. Based on reasonable assumptions, the centers of these massive halos are gaining stellar mass at a rate of 1--9\\% per Gyr on average. Compared to the merger rate for the overall population of luminous red galaxies, we find that the rate is 2--9 times greater when restricted to these dense environments. Our results imply that the massive end of the galaxy population continues to evolve hierarchically at a measurable level, and that the centers of massive groups are the preferred environment for the merger-driven assembly of massive ellipticals. ", "introduction": "Understanding the formation of the most-massive galaxies (${\\rm M}_{\\rm star}>10^{11}{\\rm M}_{\\sun}$) remains an important challenge in astrophysics. The tip of the stellar mass function is dominated by elliptical galaxies with intrinsically spheroidal mass distributions that are supported by anisotropic stellar motions \\citep{kormendy96,burstein97}. Numerical simulations have long demonstrated that ``major'' mergers between smaller galaxies of comparable mass could produce the observed shapes and dynamics of ellipticals \\citep{toomre77,barnes96,naab03,cox06}. Moreover, massive ellipticals are found in greater abundance in high-density structures like large groups and clusters of galaxies \\citep[e.g.,][]{dressler80a,postman84,hashimoto99,smith05}, which naturally grow through the hierarchical merging of dark-matter halos over cosmic time as expected in the $\\Lambda$CDM cosmological model \\citep{blumenthal84,davis85,cole00}. There is, therefore, a clear expectation for galaxy-galaxy and halo-halo merging to be physically linked \\citep{maller06,hopkins06b,delucia07}. Indeed, modern galaxy formation models predict that massive ellipticals form by major dissipationless (so-called ``dry'') merging of likewise spheroidal and gas-poor progenitors \\citep{boylan06,naab06a}, that a large fraction of today's massive ellipticals had their last major merger since redshift $z=0.5$ \\citep[e.g.,][]{delucia06}, and that the most-massive systems form at the centers of large dark-matter halos \\citep{dubinski98,aragon98}. Yet, direct evidence for the major-merger assembly of massive galaxies at present times has been lacking, and finding such systems is needed to place constraints on their rates, progenitor properties, and environmental dependencies. To this end we look for close pairs of massive interacting galaxies within a complete and well-defined sample of over 5000 galaxies with $z\\leq0.12$ and ${\\rm M}_{\\rm star}\\geq5\\times10^{10}{\\rm M}_{\\sun}$, selected from galaxy groups in the Sloan Digital Sky Survey (SDSS) with dark-matter halo masses above ${\\rm M}_{\\rm halo}=2.5\\times10^{13}{\\rm M}_{\\sun}$. Ellipticals galaxies make up the bulk of the massive end of the red-sequence population with optical colors indicative of their non-star-forming and old stellar nature. Despite a quiet star-formation history over the last 6--8 billion years \\citep{bell05a}, the total stellar mass density on the red sequence has roughly doubled over this interval \\citep{bell04b,blanton06,borch06,faber07,brown07} and now accounts for more than half of the present-day budget \\citep{hogg02,bell03b}, providing strong observational evidence for the ongoing hierarchical growth of the massive galaxy population. These results were derived from red galaxy number densities over a wide range of stellar masses above and below $10^{11}{\\rm M}_{\\sun}$. Owing to the scarcity of the highest-mass galaxies, cosmic variance, and systematic uncertainties in stellar mass estimates, any increase in the number density of ${\\rm M}_{\\rm star}>10^{11}{\\rm M}_{\\sun}$ galaxies is poorly constrained, resulting in controversy over whether this population has continued to grow slowly \\citep[e.g.,][]{brown07} or has been effectively static \\citep[e.g.,][]{scarlata07}, since $z\\sim1$. Besides number density evolution, mergers of sufficiently massive galaxies could provide a more clear indication for some continued stellar mass growth in the high-mass galaxy population. The existence of a handful of massive red mergers over the redshift interval $0.110^{11} {\\rm M}_{\\sun}$ systems. Many studies have identified major-merger candidates by either close pairs \\citep{carlberg94,patton00,carlberg00,patton02,bundy04,lin04} or disturbed morphologies \\citep{lefevre00,conselice03,lotz06}, but these samples mostly contain major mergers between lower-luminosity galaxies that tend to be gas-rich spiral disks. Numerical simulations show that such dissipative merging of disk galaxies will not produce massive pressure-supported ellipticals \\citep[e.g.,][]{naab06d}. As mentioned above, only circumstantial evidence and a small number of red galaxy pairs with $z<0.9$ support the existence of mergers likely to produce massive ellipticals. Our understanding of the progenitors is therefore very limited. Here we present a thorough census of 38 massive merger pairs from SDSS, providing an order-of-magnitude increase in the number of such detections at $z<0.5$ and allowing an improved understanding of their progenitor properties. While many estimates of major merger rates are found in the literature, to date no measure of the environmental dependence of merger-driven mass growth has been attempted. In the standard cosmological model, there is a trade off between the expansion of the universe and the gravitational collapse of dark and luminous matter. Therefore, the rate at which stellar mass is assembled at the centers of the largest dark matter halos over recent cosmic history is a fundamental aspect of the ongoing formation of large-scale structure, and the rate that high-mass galaxies form by mergers as a function of halo mass constrains galaxy formation theories. Some theories predict that the mergers producing massive ellipticals occur preferentially in groups rather than in high-density cluster or low-density field environments because the smaller velocity dispersions allow more galaxy interactions \\citep{cavaliere92}; also dynamical friction is more efficient in lower-mass halos \\citep[e.g.][]{cooray05d}. Others predict that the brightest cluster galaxies (BCGs) grow by hierarchical merging (``galactic cannibalism'') at the centers of the dark-matter potential wells of large clusters \\citep{ostriker75,merritt85,dubinski98,cooray05d}. A handful of low-redshift BCGs show multiple nuclei suggesting cannibalism in the form of multiple minor mergers \\citep{lauer88}, but there are no observations of major mergers at the centers of clusters. In this paper we make use of the statistically large SDSS group catalog \\citep{yang05a,weinmann06a} to show that major mergers occur in present-day dense environments, and to explore the halo-mass dependence and central/satellite identity of merger-driven massive galaxy assembly. Throughout this paper we calculate comoving distances in the $\\Lambda$CDM concordance cosmology with $\\Omega_{\\rm m} = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and assume a Hubble constant of $H_0 = 70 $\\,km\\,s$^{-1}$\\,Mpc$^{-1}$. SDSS magnitudes are in the AB system. ", "conclusions": "\\label{sec:Disc} We find the first direct observational evidence for an important population of galaxy-galaxy mergers with total stellar masses above $10^{11} {\\rm M}_{\\sun}$ in the local universe. These objects provide an unprecedented census of the progenitor properties for the merger-driven assembly of high-mass galaxies, which we compare to recent predictions from numerical models of galaxy formation and evolution. Moreover, the existence of these mergers prove that a measurable amount of stellar mass growth continues in the massive galaxy population at present times, and we compare estimates based on this sample with other estimates in the literature. Finally, we have identified mergers restricted to reside in large SDSS groups and clusters with $z\\leq0.12$, thus allowing the first constraints on the halo-mass dependencies of recent massive merger activity. While it is well-established that massive galaxies are more common in such high-density environments, we are missing much more than 50\\% of the population with ${\\rm M}_{\\rm star}<4\\times10^{11} {\\rm M}_{\\sun}$ in the local volume, as Table \\ref{mstar_counts} shows. Therefore, we must keep this caveat in mind when interpreting the conditions for which our results hold. In an upcoming study, we are examining the role of major mergers as a function of stellar mass over the full range of environments hosting galaxies more massive than ${\\rm M}_{\\rm star}=5\\times10^{10}{\\rm M}_{\\sun}$. \\subsection{Massive Merger Progenitors: Observations Meet Theories} \\label{sec:Disc1} Establishing the luminosity dependence of elliptical (E) galaxy properties \\citep{davies83,bender88,bender92} set the stage for theories regarding the types of merger progenitors that would produce the characteristics of low and high-mass early-type galaxies (ETGs)\\footnote{The distinction between elliptical and early-type galaxies is often blurred in the literature. We consider Es to be a morphological subset of ETGs, which are concentrated and spheroid-dominated systems including Es, lenticulars (S0s), and Sa spirals. When referencing other authors we remain faithful to their choice of nomenclature.} galaxies \\citep{bender92,kormendy96,faber97}. We concentrate on modern numerical simulations and semi-analytic models that attempt to reproduce the kinematic, photometric, and structural properties observed in massive Es through major merging \\citep{naab99,naab03,khochfar03,khochfar05,naab06a,boylan06,kang07}. For this discussion we make the straight-forward assumption that the major mergers that we have identified will produce remnants that are {\\it not unlike} the ${\\rm M}_{\\rm star}>10^{11} {\\rm M}_{\\sun}$ galaxy population already in place. We can only guess at remnant properties (see Fig. \\ref{fig:cm.remn}), but in general, massive galaxies on the red-sequence are typically early-type. As we show in Figure \\ref{fig:prog.mass}, the progenitor masses are comparable for the most part, and quantitatively consistent with the LRG-LRG merger mass spectrum from \\citet{masjedi07} under the assumption that companions merge on dynamical friction time scales. $N$-body simulations \\citep[e.g.][]{naab99} have long shown that ${\\rm M}_1/{\\rm M}_2\\approx1$ are necessary to produce the lack of significant rotation observed in massive Es. Yet, a near unity mass ratio alone is not sufficient to produce the predominance of boxy and anisotropic Es found at high luminosity \\citep{naab03,naab06a}. To match the decreasing fraction of rotational support and increasing fraction of boxiness in more luminous Es, the role of gas dissipation must be significantly reduced at high masses \\citep{bender92,khochfar05,naab06d,kang07}, and recent ETG-ETG merger simulations have demonstrated this numerically \\citep{naab06a}. Figures \\ref{fig:prog.type} and \\ref{fig:cm.prog} show that 90\\% of the progenitors in this study have concentrated light profiles and red-sequence colors, both common attributes of ETGs, with little or no cold gas content. In addition, the tidal signatures of the bulk of these massive mergers (see Figs. \\ref{fig:dpairs1} \\& \\ref{fig:dpairs2}) match those of observed \\citep{bell06a} and simulated \\citep{naab06a} major dissipationless (or gas-poor) merging of ETGs. Thus, our sample represents a more than order-of-magnitude increase in the number of such known systems with $z<0.2$, and demonstrates that dissipationless merging is indeed an important channel for the formation of massive galaxies. Finally, we compare the observed high fraction of ETG-ETG mergers ($f_{\\rm ETG-ETG}=0.9$) with several semi-analytic predictions. Recall that we have looked for signs of interaction in $>200$ major pairs from a total sample of $>5000$ massive galaxies (i.e., sampM), yet only 10\\% of the 38 mergers we identify could possibly form a ${\\rm M}_{\\rm star}>10^{11}{\\rm M}_{\\sun}$ remnant by other than an ETG-ETG merger. The progenitor morphologies of this study best match the predictions of \\citet{khochfar03}, who find $f_{\\rm E-E}=0.75$ for the last major merger of $4L^{\\ast}$ remnants, independent of environment. We find much larger ETG-ETG fractions than \\citet{naab06a} who predict only 20-35\\% (also independent of environment) over the estimated mass range of our merger remnants ($11.1<\\log_{10}{({\\rm M}_{\\rm star}/{\\rm M}_{\\sun})}<11.7$), and \\citet{kang07} who predict $f_{\\rm ETG-ETG}<0.1$ for $\\log_{10}{({\\rm M}_{\\rm star}/{\\rm M}_{\\sun})}>11$. We note that these predicted progenitor morphologies for present-day Es are based on the final major mergers that could occur over a large redshift range out to $z\\sim1$, which could be different in nature to those that occur in the short time interval that we observe. Moreover, we focus on high-density environments known to have very few massive late-type (blue) galaxies \\citep{butcher78a}, which might explain the low number of ``mixed'' (early-late or elliptical-spiral) mergers that we find. Hence, for these models to be consistent with our data, either (1) the $f_{\\rm ETG-ETG}$ of present-day major mergers depends on halo mass (i.e., environment), or (2) the relative importance of major mixed mergers has decreased significantly since $z=1$. \\subsection{Estimating Stellar Mass Accretion Rates} The existence of massive dissipationless mergers at low redshift is direct observational evidence that the growth of ${\\rm M}_{\\rm star}>10^{11}{\\rm M}_{\\sun}$ galaxies continues at present times in agreement with many cosmologically-motivated simulations \\citep{khochfar05,delucia06,kaviraj07,kang07}. Moreover, even under conservative assumptions that limit the amount of companion mass that is added to massive CEN galaxies, all of our sample will still result in remnants with ${\\rm M}_{\\rm star}>10^{11}{\\rm M}_{\\sun}$. Previously, the observational evidence for recent merger-based assembly of $z\\sim0$ massive Es was limited to luminous/massive galaxy clustering statistics \\citep{masjedi06,bell06b,masjedi07} or post-merger signatures that cannot distinguish between minor and major merging; e.g., tidal shells \\citep{malin83}, fine structure \\citep{schweizer92}, faint tidal features \\citep{vandokkum05,mihos05}, or kinematic/photometric properties \\citep[e.g.,][]{kang07}. With the merger sample presented here we can quantify directly the amount of growth, occurring in dense environments, at the high-mass end of the stellar mass function. Going from the observed merger counts to an inferred merger rate is limited mostly by the uncertainty in the merger timescale ($t_{\\rm merg}$) that one assumes. Numerical models show that the time interval for two galaxies to interact and finally merge into a single remnant depends critically on the orbital parameters, progenitor mass ratios and densities, and the degree to which the merger is dissipationless. For major mergers of massive galaxies a number of different $t_{\\rm merg}$ have been put forth in the literature based on simple orbital timescale arguments. For example, \\citet{masjedi06} derived a reasonable lower limit of $t_{\\rm merg}=0.2$ Gyr for a close ($d_{1,2}=10$ kpc) pair of LRG galaxies with a velocity dispersion of $\\sigma=200$ \\kms. Naturally, bound pairs with $d_{1,2}>10$ kpc separation will take longer to merge. \\citet{bell06b} made a similar calculation for somewhat less-massive galaxies typically separated by $d_{1,2}=15$ kpc and estimated $t_{\\rm merg}=0.4$ Gyr and argued for at least a factor of two uncertainty in this time. The mergers in this study have an average projected separation of 15.5 kpc (see Fig. \\ref{fig:pair.mergVnon}), so in what follows, we adopt $t_{\\rm merg}=0.4^{+0.4}_{-0.2}$ Gyr with conservative error bars that encompass the range of uncertainties discussed in the literature. Here, we compute the rate of stellar mass accretion by major merging onto massive galaxies in large groups. First, we find that the total mass accreted onto the centers of the $N_{\\rm CEN}=845$ halos that we study is $\\sum{ f{\\rm M}_{{\\rm s},i}}=3.9(3.5)\\times10^{12} {\\rm M}_{\\sun}$, if we include (exclude) the four SAT-SAT mergers at their host's dynamical center (see \\S \\ref{sec:cenmerging}). ${\\rm M}_{{\\rm s},i}$ is the stellar mass of the secondary (SAT) galaxy in the $i^{\\rm th}$ CEN-SAT merger, and $f$ is the fraction of ${\\rm M}_{{\\rm s},i}$ that winds up as part of the CEN galaxy. The rate of stellar mass buildup per massive CEN galaxy is therefore \\begin{equation} \\dot{{\\rm M}}_{\\rm CEN} = \\frac{\\sum{ f{\\rm M}_{{\\rm s},i} }}{N_{\\rm CEN}} \\times \\frac{1}{t_{\\rm merg}} , \\label{eq:4} \\end{equation} or between $1.0^{+1.0}_{-0.5}\\times10^{10}{\\rm M}_{\\sun}{\\rm Gyr}^{-1}$ and $1.2^{+1.1}_{-0.6}\\times10^{10}{\\rm M}_{\\sun}{\\rm Gyr}^{-1}$, depending on which sample of CEN-SAT mergers that we consider. The lopsided error bars result from the range of accretion rates for $t_{\\rm merg}=0.4^{+0.4}_{-0.2}$ Gyr, as described above. If we divide all of these accretion rates by $2.69\\times10^{11} {\\rm M}_{\\sun}$, the average stellar mass of the 845 CEN galaxies in this study, we find that each CEN is growing by 1--9\\% per Gyr. Finally, these values can be decreased by assuming $f<1$ in (\\ref{eq:4}), but as we discuss in \\S \\ref{sec:cmplane}, $f=0.5$ represents a likely lower limit. Likewise, the total stellar mass accreted onto all galaxies in sampM is $\\sum{ f{\\rm M}_{{\\rm s},i}}+\\sum{ {\\rm M}_{{\\rm s},j}}=5.1\\times10^{12} {\\rm M}_{\\sun}$, where ${\\rm M}_{{\\rm s},j}$ is the mass of the secondary (SAT) galaxy in the $j^{\\rm th}$ SAT-SAT merger. Therefore, the growth per ${\\rm M}_{\\rm star}\\geq5\\times10^{10} {\\rm M}_{\\sun}$ galaxy in high-mass groups is \\begin{equation} \\dot{{\\rm M}}_{(\\geq5\\times10^{10} {\\rm M}_{\\sun})} = \\frac{\\sum{ f{\\rm M}_{{\\rm s},j} } + \\sum{ {\\rm M}_{{\\rm s},j}} }{ (N_{\\rm CEN}+N_{\\rm SAT}-N_{\\rm s,sampM})} \\times \\frac{1}{t_{\\rm merg}} , \\end{equation} where $N_{\\rm s,sampM}=12$ is the number of secondary SAT galaxies in sampM that are involved in major mergers and must be subtracted to avoid double counting. We find $\\dot{{\\rm M}}_{(\\geq5\\times10^{10} {\\rm M}_{\\sun})}=2.4^{+2.4}_{-1.2}\\times10^{9}{\\rm M}_{\\sun}{\\rm Gyr}^{-1}$; if we assume $f=0.5$ for CEN-SAT mergers only we find $\\dot{{\\rm M}}_{(\\geq5\\times10^{10} {\\rm M}_{\\sun})}=1.6^{+1.5}_{-0.7}\\times10^{9}{\\rm M}_{\\sun}{\\rm Gyr}^{-1}$. Given that the average stellar mass of sampM galaxies is $1.04\\times10^{11} {\\rm M}_{\\sun}$, we find that every massive galaxy is growing by 1--5\\% per Gyr. Even though SAT-SAT mergers may occur as frequently as CEN-SAT mergers in these massive groups, the centers are where much of the mass growth takes place. It is clear from Figure \\ref{fig:prog.mass} that mostly only ${\\rm M}_{\\rm star}>10^{11} {\\rm M}_{\\sun}$ galaxies build up in mass by major mergers in groups with ${\\rm M}_{\\rm halo}>2.5\\times10^{13}{\\rm M}_{\\sun}$. In contrast, we find few mergers among the $5\\times 10^{10}<{\\rm M}_{\\rm star}<10^{11} {\\rm M}_{\\sun}$ galaxies in these high-mass groups, which make up the bulk (60\\%) of sampM. This suggests that if major merging is playing an important role in the strong mass growth observed on the red sequence below M* \\citep{bell04b,blanton06,borch06,faber07,brown07}, it is occurring in lower-mass groups than we study here. Rather than mass growth rates we can use the same line of reasoning to estimate massive galaxy-galaxy merging rates of $(21+4)/845/t_{\\rm merg}=0.074^{+0.074}_{-0.037}{\\rm Gyr}^{-1}$ for CEN-SAT and $(38+7)/(845+4531-12)/t_{\\rm merg}=0.021^{+0.021}_{-0.011}{\\rm Gyr}^{-1}$ for all galaxies in sampM. For these estimates we included the seven additional major mergers (4 CEN, 3 SAT) we identified by their highly-disturbed appearance. \\citet{masjedi06} found a strict upper limit to the LRG-LRG rate of only $0.006{\\rm Gyr}^{-1}$. We estimate that LRGs have a stellar mass range of $11.4<\\log_{10}{({\\rm M}_{\\rm star}/{\\rm M}_{\\sun})}<12.0$, based on typical red-sequence colors and luminosities between $4L^{\\ast}$ and $25L^{\\ast}$. Within these mass limits, we find a merger rate of $5/462/t_{\\rm merg}=0.027^{+0.027}_{-0.014}{\\rm Gyr}^{-1}$ on the red sequence, or 2--9 times the LRG-LRG rate. In Table \\ref{mstar_counts}, we show that the high-mass groups that we study contain $>70\\%$ of the very-massive, red galaxy population in the $z\\leq0.12$ volume of DR2, with the vast majority being CENs. Yet, the same group selection contains only 30\\% of the population of $11.4<\\log_{10}{({\\rm M}_{\\rm star}/{\\rm M}_{\\sun})}<11.6$ systems. These numbers show that a significant portion of the local counterparts to LRGs are found in groups with ${\\rm M}_{\\rm halo}<2.5\\times10^{13}{\\rm M}_{\\sun}$. Therefore, we conclude that LRG-LRG merging occurs more frequently in the more massive groups." }, "0710/0710.1542_arXiv.txt": { "abstract": "{The initial-final mass relationship of white dwarfs, which is poorly constrained, is of paramount importance for different aspects in modern astrophysics. From an observational perspective, most of the studies up to now have been done using white dwarfs in open clusters.} {In order to improve the initial-final mass relationship we explore the possibility of deriving a semi-empirical relation studying white dwarfs in common proper motion pairs. If these systems are comprised of a white dwarf and a FGK star, the total age and the metallicity of the progenitor of the white dwarf can be inferred from the detailed analysis of the companion.} {We have performed an exhaustive search of common proper motion pairs containing a DA white dwarf and a FGK star using the available literature and crossing the SIMBAD database with the Villanova White Dwarf Catalog. We have acquired long-slit spectra of the white dwarf members of the selected common proper motion pairs, as well as high resolution spectra of their companions. From these observations, a full analysis of the two members of each common proper motion pair leads to the initial and final masses of the white dwarfs.} {These observations have allowed us to provide updated information for the white dwarfs, since some of them were misclassified. In the case of the DA white dwarfs, their atmospheric parameters, masses, and cooling times, have been derived using appropriate white dwarf models and cooling sequences. From a detailed analysis of the FGK stars spectra we have inferred the metallicity. Then, using either isochrones or X-ray luminosities we have obtained the main-sequence lifetime of the progenitors, and subsequently their initial masses.} {This work is the first one in using common proper motion pairs to improve the initial-final mass relationship, and has also allowed to cover the poorly explored low-mass domain. As in the case of studies based on white dwarfs in open clusters, the distribution of the semi-empirical data presents a large scatter, which is higher than the expected uncertainties in the derived values. This suggests that the initial-final mass relationship may not be a single-valued function.} ", "introduction": "White dwarfs are the final remnants of low- and intermediate-mass stars. About 95\\% of main-sequence stars will end their evolutionary pathways as white dwarfs and, hence, the study of the white dwarf population provides details about the late stages of the life of the vast majority of stars. Since white dwarfs are long-lived objects, they also constitute useful objects to study the structure and evolution of our Galaxy (Liebert et al.~2005a; Isern et al. 2001). For instance, the initial-final mass relationship (IFMR), which connects the properties of a white dwarf with those of its main-sequence progenitor, is of paramount importance for different aspects in modern astrophysics. It is required as an input for determining the ages of globular clusters and their distances, for studying the chemical evolution of galaxies, and also to understand the properties of the Galactic population of white dwarfs. Despite its relevance, this relationship is still poorly constrained, both from the theoretical and the observational points of view. The first attempt to empirically determine the initial-final mass relationship was undertaken by \\cite{wei77}, who also provides a recent review on this subject (Weidemann 2000). It is still not clear how this function depends on the mass and metallicity of the progenitor, its angular momentum, or the presence of a strong magnetic field. The total age of a white dwarf can be expressed as the sum of its cooling time and the main-sequence lifetime of its progenitor. The latter depends on the metallicity of the progenitor of the white dwarf, but it cannot be determined from observations of single white dwarfs. This is because white dwarfs have such strong surface gravities that gravitational settling operates very efficiently in their atmospheres, and any information about their progenitors (e.g. metallicity) is lost in the very early evolutionary stages of the cooling track. Moreover, the evolution during the AGB phase of the progenitors is essential in determining the size and composition of the atmospheres of the resulting white dwarfs, since the burning processes that take place in H and He shells determine their respective thicknesses and their detailed chemical compositions, which are crucial ingredients for determining the evolutionary cooling times. A promising approach to circumvent the problem, and also to directly test the initial-final mass relationship, is to study white dwarfs for which external constraints are available. This is the case of white dwarfs in open and globular clusters (Ferrario et al.~2005, Dobbie et al.~2006) or in non-interacting binaries, for instance, common proper motion pairs (Wegner 1973, Oswalt et al.~1988). Focusing on the latter, it is sound to assume that the members of a common proper motion pair were born simultaneously and with the same chemical composition. Since the components are well separated (100 to 1000 AU), mass exchange between them is unlikely and it can be considered that they have evolved as isolated stars. Thus, important information of the white dwarf, such as its total age or the metallicity of the progenitor, can be inferred from the study of the companion. In particular, if the companion is an F, G or K type star the metallicity can be derived with high accuracy from detailed spectral analysis. On the other hand, the age can be obtained using different methods. In particular, we will use stellar isochrones when the star is moderately evolved, or the X-ray luminosity if the star is very close to the ZAMS. The purpose of this work is to present our spectroscopic analysis of both members of some common proper motion pairs containing a white dwarf, and the semi-empirical intial-final mass relationship that we have derived from this study. The paper is organized as follows. In \\S 2 we present the observations done so far and describe the data reduction. Section 3 is devoted to discuss the classification and the analysis of the observed white dwarfs, whereas in \\S 4 we present the analysis of the companions. This is followed by \\S 5 where we present our main results and finally in \\S 6 we elaborate our conclusions. ", "conclusions": "We have studied a sample of common proper motion pairs comprised of a white dwarf and a FGK star. We have performed high signal-to-noise low resolution spectroscopy of the white dwarf members, which led us to carry out a full analysis of their spectra and to make a re-classification when necessary. From the fit of their spectra to white dwarf models we have derived their atmospheric parameters. Then, using different cooling sequences --- namely those of \\cite{sal00} and \\cite{fon01} --- their masses and cooling times were obtained. Simultaneously, we have performed independent high resolution spectroscopic observations of their companions. Using the available photometry we have obtained their effective temperatures. Then, from a detailed analysis of their spectra and using either isochrones or X-ray luminosities, we have derived their metallicities and ages (i.e., the metallicities of the progenitors of the white dwarfs and their total ages). These observations allowed us to obtain the initial and final masses of six white dwarfs in common proper motion pairs, four of them corresponding to initial masses below $2\\,\\rm M_{\\sun}$, a range which has not been previously covered by the open cluster data. Our semi-empirical relation shows significant scatter, compatible with the results obtained by \\cite{fer05} and \\cite{dob06}, which are mainly based on open cluster data. However, the dispersion of the results is higher than the error bars, which leaves some open questions that should be studied in detail (e.g., rotation or magnetic fields). We have shown that common proper motion pairs containing white dwarfs can be useful to improve the initial-final mass relationship, since they cover a wide range of ages, masses and metallicities, and they are also representative of the disk white dwarf population. We have seen that the accuracy in the total ages depends almost exclusively on the evolutionary state of the low-mass companions. Such relative accuracy becomes poor when the star is close to the ZAMS. However, this limitation may not be critical to many common proper motion pairs. Planned deep surveys like GAIA, LSST or the Alhambra Survey will discover thousands of new white dwarfs, some of them belonging to wide binaries. In the meantime, our most immediate priority is to further extend the sample of wide binaries valid for this study. We are working in the search for more wide binaries of our interest in the NLTT catalog (Gould \\& Chanam\\'e 2004) and also in the LSPM-north catalog (L\\'epine \\& Bongiorno 2007). Detailed study of the current and future common proper motion pairs of this type should help to explain the scatter in the semi-empirical initial-final mass relationship and to discern whether this is a single-valued function. If consistency between observations and theoretical calculations is found, this would have a strong impact on stellar astrophysics, since this relationship is used in many different areas, such as chemical evolution of galaxies, the determination of supernova rates or star formation and feedback processes in galaxies." }, "0710/0710.4151_arXiv.txt": { "abstract": "The discovery of dark energy (DE) as the physical cause for the accelerated expansion of the Universe is the most remarkable experimental finding of modern cosmology. However, it leads to insurmountable theoretical difficulties from the point of view of fundamental physics. Inflation, on the other hand, constitutes another crucial ingredient, which seems necessary to solve other cosmological conundrums and provides the primeval quantum seeds for structure formation. One may wonder if there is any deep relationship between these two paradigms. In this work, we suggest that the existence of the DE in the present Universe could be linked to the quantum field theoretical mechanism that may have triggered primordial inflation in the early Universe. This mechanism, based on quantum conformal symmetry, induces a logarithmic, asymptotically-free, running of the gravitational coupling. If this evolution persists in the present Universe, and if matter is conserved, the general covariance of Einstein's equations demands the existence of dynamical DE in the form of a running cosmological term, $\\CC$, whose variation follows a power law of the redshift. ", "introduction": "Modern Cosmology incorporates the notion of dark energy (DE) as an experimental fact that accounts for the physical explanation of the observed accelerated expansion\\,\\cite{SNe,WMAP3Y}. Although the nature of the DE is not known, one persistent possibility is the 90-years-old cosmological constant (CC) term, $\\CC$, in Einstein's equations. In recent times, one is tempted to supersede this hypothesis with another, radically different, one: viz. a slowly evolving scalar field $\\phi$ (``quintessence'') whose potential, $V(\\phi)\\gtrsim 0$, could explain the present value of the DE and whose equation of state (EOS) parameter $\\omega_{\\phi}= p_{\\phi}/\\rho_{\\phi}\\simeq -1+\\dot\\phi^2/V(\\phi)$ is only slightly larger than $-1$ (hence insuring a negative pressure mimicking the $\\CC$ case)\\,\\cite{weinberg}. The advantage to think this way is that the DE can then be a dynamical quantity taking different values throughout the history of the Universe. However, this possibility can not explain why the DE is entirely due to such an \\textit{ad hoc} scalar field and why the contributions to the vacuum energy from the other fields (e.g. the electroweak Standard Model ones) must not be considered. In short, it does not seem to be such a wonderful idea to invent the field $\\phi$ and simply replace $\\rL=\\CC/8\\pi\\,G$ (the energy density associated to $\\CC$, where $G$ is Newton's constant) with $\\rho_{\\phi}\\simeq V(\\phi)$. One has to explain, too, why the various contributions (including the additional one $V(\\phi)$!) must conspire to generate the tiny value of the DE density at present -- the ``old CC problem''\\,\\cite{weinberg}. While we cannot solve this problem at this stage, the dynamical nature of the DE makes allowance for this possibility. Furthermore, since there is no obvious gain in the quintessence idea, we stick to the CC approach, although we extend it to include the possibility of a dynamical (``running'') $\\CC$ term\\,\\cite{JHEPCC1,SSRev}. The obvious question now is: where this dynamics could come from? One possibility is that it could originate from the fundamental mechanism of inflation\\,\\cite{inflation}, which presumably took place in the very early Universe and could have left some loose end or remnant -- kind of ``fossil'' -- in our late Universe, which we don't know where to fit in now. However, what mechanism of inflation could possibly do that? There is in principle a class of distinct possibilities, in particular see \\cite{Sahni98,Mongan01}, but our very source of inspiration here is the quantum theory of the conformal factor, which was extensively developed in\\,\\cite{Conformal1}. For a recent discussion, see e.g. \\cite{Conformal2,Conformal3} and references therein. More specifically, we start from the idea of ``tempered anomaly-induced inflation'', which was first proposed in \\cite{shocom,DESY} (see also \\cite{PST}). It leads essentially to a modified form of the original Starobinsky model\\,\\cite{Starobinsky}. In the present paper, we push forward the possibility that the mechanism that successively caused, stabilized, slowed down (``tempered'') and extinguished the fast period of inflation in our remote past could have left an indelible imprint in the current Universe, namely a very mild (logarithmically) running Newton's coupling $G$. We show that, if matter is covariantly conserved, this necessarily implies an effective renormalization group (RG) running of the ``cosmological constant'' energy density, $\\rL=\\rL(a)$, which takes the form of a cubic law of $a^{-1}=1+z$ during the matter dominated epoch ($a$ being the scale factor and $z$ the cosmological reshift). ", "conclusions": "In this work we have suggested that the presence of dynamical dark energy (DE) in the current Universe is actually a consistency demand of Einstein equations under the two assumptions of: i) matter conservation, and ii) the existence of a period of primordial inflation in the early Universe, especially when realized as ``tempered anomaly-induced inflation''. Based essentially on the previous works\\,\\cite{shocom,DESY,PST} and on the general setting of the quantum theory of the conformal factor\\,\\cite{Conformal1,Conformal2}, we have found that if the inflationary mechanism is caused by quantum effects on the effective action of conformal quantum field theory in curved space-time, then the gravitational coupling $G$ becomes a running quantity of the scale factor, $G(a)=G_0/(1-\\f\\ln a)$, $\\f$ being the coefficient of the $\\beta$-function for the conformal Newton's coupling. The effect of this coupling on the inflationary dynamics is to efficiently ``temper'' the regime of stable inflation presumably into the FLRW regime. The rigorous high energy calculation of $\\f$ in QFT in curved space-time shows that both fermions and bosons produce non-negative contributions ($\\f\\geq 0$). As a consequence, $G$ becomes an asymptotically-free coupling of the scale factor. Intriguingly enough, we have suggested the possibility that this running might persist in the present Universe and, if so, it could provide a \\textit{raison d'\\^etre} for the existence of the (dynamical) DE, which would appear in the form of running cosmological vacuum energy $\\rL$. In fact, the logarithmic evolution of $G$ induces a power-law running of $\\rL$, which is essentially driven by the soft-decoupling terms $\\sim H^2\\,M_i^2$ (hence by the heaviest particle masses). The result is a Universe effectively filled with a mildly-dynamical DE, which can be perfectly consistent with the present observations. To summarize, from the point of view of the ``RG-cosmology'' under consideration, the current Universe appears as FLRW-like while still carrying some slight imprints of important physical processes that determined the early stages of the cosmic evolution. Most conspicuously, the smooth dynamics of $G$ and $\\rL$ can be thought of as ``living fossils'' left out of the quantum field theoretical mechanism that triggered primordial inflation. Remarkably, this framework fits with previous attempts to describe the renormalization group running of the cosmological term\\,\\cite{JHEPCC1,RGTypeIa,IRGA03,SSS1,croat,SS12,LXCDM12,Bilic07} and could provide an attractive link between all stages of the cosmic evolution. It is reassuring to find that there is a large class of RG models behaving effectively the same way. Differences between them could probably be resolved at the level of finer tests, such as those based on cosmological perturbations and structure formation. For example, in references\\,\\cite{FSS1,GOPS} it is shown that the study of cosmological perturbations within models of running cosmological constant puts a limit on the amount of running, which is more or less stringent depending on the peculiarities of the model. Similarly, a particular study of perturbations would be required in the present framework (which includes the variation of both $\\Lambda$ and $G$) to assess the implications on the parameter $\\f$. This study is beyond the scope of the present work. \\vspace{1cm} \\textit{Acknowledgements}. I am very grateful to Ilya Shapiro for discussions on different aspects of this work and for the fruitful collaboration maintained on the cosmological constant ptoblem over the years. I thank also Ana Pelinson for interesting discussions in the early stages of this work. The author has been supported in part by MECYT and FEDER under project 2004-04582-C02-01, and also by DURSI Generalitat de Catalunya under 2005SGR00564 and the Brazilian agency FAPEMIG. I am thankful for the warm hospitality at the Dept. of Physics of the Univ. Federal de Juiz de Fora, where part of this work was carried out." }, "0710/0710.1318_arXiv.txt": { "abstract": "We present preliminary results on the calculation of synthetic spectra obtained with the stellar model atmospheres developed by Cardona, Crivellari, and Simonneau. These new models have been used as input within the {\\sc Synthe} series of codes developed by Kurucz. As a first step we have tested if {\\sc Synthe} is able to handle these models which go down to $\\log{\\tau_{\\rm Ross}}= -13$. We have successfully calculated a synthetic solar spectrum in the wavelength region 2000--4500~\\AA\\ at high resolution ($R=522\\,000$). Within this initial test we have found that layers at optical depths with $\\log{\\tau_{\\rm Ross}} < -7$ significantly affect the mid-UV properties of a synthetic spectrum computed from a solar model. We anticipate that these new extended models will be a valuable tool for the analysis of UV stellar light arising from the outermost layers of the atmospheres. \\end {abstract} ", "introduction": "\\label{sec:1} A set of spectral energy distributions (SEDs) is a very useful tool to analyze stellar spectra and the integrated spectral properties of stellar systems (via some evolutionary population synthesis code, see e.g. Buzzoni 1995). Observational atlases of stellar SEDs generally lack of homogeneous and complete coverage of the main stellar parameters ($T_{\\rm eff}, \\log{g}$ and [M/H]), therefore many of the recent population analyses rely on results of stellar atmosphere modelling, a practise that has been eased with the development of faster computers and sophisticated computational codes. In order to calculate a theoretical stellar SED it is necessary to have a model atmosphere which describes the physical quantities at different depths and, ideally, a complete set of opacities that account for the absorption of the radiation passing throughout the atmosphere. During the last decades several groups have developed computational codes capable to calculate model atmospheres and spectra at high resolution, some of whom have allowed the public use of their codes, among others, the codes {\\sc Atlas9}, {\\sc Atlas12} and {\\sc Synthe} built by Kurucz\\footnote{http://kurucz.harvard.edu/} (1993a,b) and {\\sc Tlusty} and {\\sc Synspec} constructed by Hubeny \\& Lanz\\footnote{http://nova.astro.umd.edu/} (1992). In particular, {\\sc Atlas9} and {\\sc Synthe} codes have been used by our group to calculate the UVBLUE grid of theoretical SEDs and to investigate its potential for stellar and populations studies in the ultraviolet (UV) wavelength interval (see Rodriguez-Merino et al. 2005 for more details). The UV wavelength range has historically been challenging in many branches of modern astrophysics since the observed stellar spectra are not well reproduced by predictions of theoretical models. One possible reason is the missing opacity problem (see Holweger 1970; Gustafsson et al. 1975), but another probable reason is actually that most of the model atmospheres do not provide the atmosphere structure near the stellar surface, where most of the UV radiation emerges. Therefore, it is crucial to calculate new models which describe the outermost layers of the stellar atmospheres. In this work we briefly describe the structure of a new model atmosphere for the Sun which incorporates layers down to $\\log{\\tau_{\\rm Ross}}=-13$. That is, optical depths more than five orders of magnitude thinner compared to classical models currently in use. This model has been couplet to {\\sc Synthe} codes for testing their compatibility and exploring the effects of such layers on the UV flux. ", "conclusions": "\\label{sec:4} The main result of this work is that {\\sc Synthe} series of codes is capable of treating the CCS model atmospheres, which reach very low values of $\\tau_{\\rm Ross}$. The analysis of the effects of extending the atmosphere indicates that at mid-UV wavelengths the effects are significant while negligible in the blue. The following steps are to extend the analysis to models with atmospheric parameters different of the Sun, to complement the opacity (both continuous and of lines) as well as to introduce more chemical species. We are in the process of including convection for intermediate and cool star models. A detailed comparison at high resolution with an observed solar atlas (Kurucz et al. 1984) is also underway." }, "0710/0710.0128_arXiv.txt": { "abstract": "% Preliminary VLBA polarisation results on 6 ``blazars'' from 6.5-cm to 7-mm are presented here. Observing at several different wavelengths, separated by short and long intervals, enabled reliable information about the magnetic (B) field structure to be obtained and for the effect of Faraday Rotation to be determined and corrected. For all sources the magnitude of the core Rotation Measure (RM) derived from the shorter wavelength data was greater than that derived from the longer wavelength data, consistent with a higher electron density and/or B-field strength closer to the central engine. A transverse RM gradient was detected in the jet of 0954+658, providing evidence for the presence of a helical B-field surrounding the jet. The RM in the core region of 2200+420 (BL Lac) displays sign changes in different wavelength intervals (on different spatial scales); we suggest an explanation for this in terms of modest bends in a helical B-field surrounding the jet. ", "introduction": "The Faraday effect causes a rotation of the plane of linear polarisation, described by: $\\Delta\\chi = RM \\lambda^2$, with the rotation measure (RM) determined by the integral of the electron density and the dot product of the magnetic (B) field and the path length along the line of sight (LoS). A positive/negative RM indicates that the LoS B-field is pointing towards/away from the observer. $\\Delta\\chi$ is the change in the electric vector position angle (EVPA). Previous results indicated the presence of different RM signs in the core regions of 6 blazars in different wavelength intervals (O'Sullivan \\& Gabuzda 2006). This has two main possible origins: (1) the LoS B-field changes with distance from the centre of activity, or (2) since the previous observations were not simultaneous, it could be due to an intrinsic change in the overall jet B-field structure between observing epochs. Our new 8 wavelength observations are designed to test these possibilities. ", "conclusions": "Our results for 2200+420 confirm the presence of an RM sign reversal in the core region. Since the dominant jet B-field is transverse to the jet and remains transverse while the jet bends, we will suppose a helical B-field surrounds the jet. The observed RM sign reversal can be explained by a slight bend of the jet, due, for example, to a collision with material in the parent galaxy or some instabilities inherent in the jet itself. (A longitudinal jet B-field with a change in the angle to the line of sight could also cause a RM sign reversal, but this does not correspond to the observed B-field.) A side-on view of a helical B-field (Fig.1 Top right) will have a RM that will be equally strong on both sides of the jet, hence, a zero net RM will be observed for an unresolved jet. This would occur when the source is viewed at $1/\\Gamma$ in the observer's frame. For a tail-on view of a helical B-field (ie. $\\theta > 1/\\Gamma$) (Fig.1 Middle), the dominant RM will be from the bottom half of the jet and a negative RM will be observed because the dominant LoS B-field will be pointing away from us. (Assuming the jet is not fully resolved in the transverse direction.) Conversely, for a head-on view of a helical B-field (ie. $\\theta < 1/\\Gamma$) (Fig.1 Bottom), a positive RM will be observed. Therefore, regions with different RM signs in the jets of AGN can be explained within a helical B-field model as places where the jet is observed at angles greater than or less than $~1/\\Gamma$, due to bends in the jet. Since VLBI resolution is usually not sufficient to completely resolve the true optically thick core, the VLBI ``core'' consists of emission from the true core and some of the optically thin inner-jet. So if bends occur on scales smaller than the observed VLBI core, ``core'' RMs with different signs could be derived from observations at different wavelengths (ie. probing different scales of the inner-jet). In our future work, we will attempt to reconstruct the 3D path of the jet through space using the combined information from the observed distributions of the total intensity, linear polarisation, spectral index and rotation measure." }, "0710/0710.5682_arXiv.txt": { "abstract": "{ {\\it Context: } One of the most debated issues about sub-mJy radio sources, which are responsible for the steepening of the 1.4 GHz source counts, is the origin of their radio emission. Particularly interesting, from this point of view, is the possibility of combining radio spectral index information with other observational properties to assess whether the sources are triggered by star formation or nuclear activity. {\\it Aims:} The aim of this work is to study the optical and near infrared properties of a complete sample of 131 radio sources with $S>0.4$ mJy, observed at both 1.4 and 5 GHz as part of the ATESP radio survey. The availability of multi--wavelength radio and optical information is exploited to infer the physical properties of the faint radio population. {\\it Methods:} We use deep multi--colour (UBVRIJK) images, mostly taken in the framework of the ESO \\emph{Deep Public Survey}, to optically identify and derive photometric redshifts for the ATESP radio sources. Deep optical coverage and extensive colour information are available for 3/4 of the region covered by the radio sample. Typical depths of the images are $U\\sim 25$, $B\\sim 26$, $V\\sim 25.4$, $R\\sim 25.5$, $I\\sim 24.3$, $19.5\\leq K_{s}\\leq 20.2$, $J\\leq 22.2$. We also add shallower optical imaging and spectroscopy obtained previously in order to perform a global analysis of the radio sample. {\\it Results:} Optical/near infrared counterparts are found for $\\sim 78\\%$ (66/85) of the radio sources in the region covered by the deep multi--colour imaging, and for 56 of these reliable estimates of the redshift and type are derived. We find that many of the sources with flat radio spectra are characterised by high radio--to--optical ratios ($R>1000$), typical of classical powerful radio galaxies and quasars. Flat--spectrum sources with low $R$ values are preferentially identified with early type galaxies, where the radio emission is most probably triggered by low--luminosity active galactic nuclei. Considering both early type galaxies and quasars as sources with an active nucleus, such sources largely dominate our sample (78\\%). Flat--spectrum sources associated with early type galaxies are quite compact ($d<10-30$ kpc), suggesting core-dominated radio emission. ", "introduction": "\\label{sec:introduction} The faint (sub-mJy) radio population consists of a mixture of different classes of objects. Since the early seventies it has been known that the strongest sources are almost exclusively associated with either active galactic nuclei (AGNs) or giant ellipticals, the latter of which are also known as radio galaxies (99\\% above 60 mJy, \\citealt{Windhorst90}). More recent work on mJy and sub-mJy sources has revealed that faint sources are also found to be associated with normal elliptical, spiral and star-forming galaxies, with the early type galaxies being the dominant component \\citep{Gruppioni1999, Georgakakis1999, Magliocchetti2000, Prandoni2001b, Afonso2006}, while at $\\mu$Jy levels star-forming galaxies prevail (see e.g. \\citealt{Richards1999}). In spite of the progress made in our understanding of the faint radio population, many questions remain open. For example, the relative fractions of the different types of objects are still quite uncertain, and our knowledge of their dependence on limiting flux density is still incomplete. The reason is, of course, that very little is known about the faint ends of the various luminosity functions, and even less is known about the cosmological evolution of different kinds of objects. This uncertainty is due to the incompleteness of optical identification and spectroscopy, since faint radio sources usually have very faint optical counterparts. Clearly {\\it very}\\, deep ($I\\apprge 25$) optical imaging and spectroscopy, for reasonably large deep radio samples, are critical if one wants to investigate these radio source populations. Since the radio emission comes from different types of objects an important question is what are the physical processes that trigger this emission. It is natural to assume that in the case of star-forming galaxies the emission traces the history of galaxy formation and subsequent evolution by merging and interaction, while the emission in AGNs will reflect black hole accretion history. To make matters more complicated, both processes may be present at the same time. Although research in this field proceeds slowly due to very time--consuming spectroscopy much progress has been made in recent years thanks to strong improvement in the photometric redshift technique. Several multi--colour/multi--object spectroscopy surveys overlapping deep radio fields have recently been undertaken, including the Phoenix Deep Survey \\citep{Hopkins1998,Georgakakis1999,Afonso2006} and the Australia Telescope ESO Slice Project (ATESP) survey \\citep{Prandoni2000a,Prandoni2000b,Prandoni2001b}. In other cases, deep multi--colour/multi--wavelength surveys have been complemented by deep radio observations (see e.g. the VLA--VIRMOS, \\citealt{Bondi2003}; and the COSMOS, \\citealt{Schinnerer2006}). Multi--frequency radio observations are also important in measuring the radio spectral index, which may help to constrain the origin of the radio emission in the faint radio sources. This approach is especially meaningful when high resolution radio images are available and radio source structures can be inferred. However, multi--frequency radio information is available for very few, and small, sub-mJy radio samples. The largest sample with multi--frequency radio coverage available so far is a complete sample of 131 radio sources with $S>0.4$ mJy, extracted from a square degree region observed at both 1.4 and 5 GHz as part of the ATESP radio survey \\citep{Prandoni2000a,Prandoni2000b,Prandoni2006}. The $1.4-5$~GHz radio spectral index analysis of the ATESP radio sources was presented in the first paper of this series (\\citealt{Prandoni2006}, hereafter Paper I). We found a flattening of the radio spectra with decreasing radio flux density. At mJy levels most sources have steep spectra ($\\alpha \\sim -0.7$, assuming $S\\sim \\nu^{\\alpha}$), typical of synchrotron radiation, while at sub-mJy flux densities a composite population is present, with up to $\\sim 60\\%$ of the sources showing flat ($\\alpha > -0.5$) spectra and a significant fraction ($\\sim 30\\%$) of inverted-spectrum ($\\alpha>0$) sources. This flattening at sub-mJy fluxes confirms previous results based on smaller samples (\\citealt{Donnelly1987,Gruppioni1997,Ciliegi2003}). Flat spectra in radio sources usually indicate the presence of a self-absorbed nuclear core, but they can also be produced on larger scales by thermal emission from stars. It is possible to combine the spectral index information with other observational properties and infer the nature of the faint radio population. This is especially important with respect to the class of flat/inverted--spectrum sources as it permits us to study the physical processes that trigger the radio emission in those sources. This kind of analysis needs information about the redshifts and types of the galaxies hosting the radio sources. A detailed radio/optical study of the sample above is possible, thanks to the extensive optical/infrared coverage mostly obtained in the ESO \\emph{Deep Public Survey} (DPS, \\citealt{Mignano2007,Olsen2006}). We give a brief discussion of all the data collected so far in Sect.~\\ref{sec:datacoverage}, followed by a more detailed analysis of the DPS optical data in Sect.~\\ref{sec:dpsanalysis}, where we derive the UBVRI colour catalogue and photometric redshifts for the DPS galaxies in the region covered by the ATESP survey, assessing the reliability of the photometric redshifts themselves. In Sects.~\\ref{sec:optid} and \\ref{sec:radiozphot}, respectively, we use the DPS UBVRIJK optical data to identify the ATESP radio sources and to derive photometric redshifts. A radio/optical analysis of the optically identified radio sources is presented in Sect.~\\ref{sec:comp}, while in Sect.~\\ref{sec:nature} we discuss the nature of the mJy and sub--mJy population on the basis of all the radio and optical data available to the ATESP sample. The main results are briefly summarised in Sect.~\\ref{sec:summary}.\\\\ Throughout this paper we use the $\\Lambda$CDM model, with $H_0=70$, $\\Omega_m=0.3$ and $\\Omega_{\\Lambda}=0.7$. ", "conclusions": "\\label{sec:summary} In this paper we have discussed the nature of the faint, sub-mJy, radio population, using a sample of 131 radio sources that were observed at 1.4 and 5 GHz with the ATCA (the ATESP--DEEP1 sample). A smaller sample of 85 radio sources is covered by deep multi--colour images. These were optically identified down to very faint magnitudes, which was possible thanks to the availability of very deep multi--colour optical material (in U, B, V, R, I, and sometimes J and K bands). The high percentage of identifications ($\\sim 78\\%$) makes this a sample that is well suited for follow up studies concerning the composition of the sub-mJy population and, in general, the cosmological evolution of the various classes of objects associated with faint radio sources. We summarise our main results here. \\begin{itemize} \\item For 85\\% of the identification sample we succeeded in deriving reliable photometric redshifts, based on the available accurate colours (UBVRIJK). \\item Based on spectral types determined either directly from spectroscopy or from the photometry (or both), we find that at the sub-mJy level the large majority of sources are associated with objects that have early type (64\\%) and AGN (14\\%) spectra; these are of course what we would normally call radio galaxies and quasars. \\item Although earlier work (based on shallower optical imaging and spectroscopy) revealed the presence of a conspicuous component of late type and star-burst objects, such objects appear to be important only at brighter magnitudes ($I<19$), and are rare at fainter magnitudes ($192$, the systematic residuals in the measured spectrum must be reduced to 3~mK. ", "introduction": "The transition period at the end of the cosmic ``Dark Ages'' is known as the epoch of reionization (EOR). During this epoch, radiation from the very first luminous sources---early stars, galaxies, and quasars---succeeded in ionizing the neutral hydrogen gas that had filled the intergalactic medium (IGM) since the recombination event following the Big Bang. Reionization marks a significant shift in the evolution of the Universe. For the first time, gravitationally-collapsed objects exerted substantial feedback on their environments through electromagnetic radiation, initiating processes that have dominated the evolution of the visible baryonic Universe ever since. The epoch of reionization, therefore, can be considered a dividing line when the relatively simple evolution of the early Universe gave way to more complicated and more interconnected processes. Although the Dark Ages are known to end when the first luminous sources ionized the neutral hydrogen in the IGM, precisely when this transition occurred remains uncertain. The best existing constraints on the timing of the reionization epoch come from two sources: the cosmic microwave background (CMB) anisotropy and absorption features in the spectra of high-redshift quasars. The amplitude of the observed temperature anisotropy in the CMB is affected by Thomson scattering due to electrons along the line of sight between the surface of last scattering and the detector, and thus, it is sensitive to the ionization history of the IGM through the electron column density. In addition, if there is sufficient optical depth to CMB photons due to free electrons in the IGM after reionization, some of the angular anisotropy in the unpolarized intensity can be converted to polarized anisotropy. This produces a peak in the polarization power spectrum at the angular scale size equivalent to the horizon at reionization with an amplitude proportional to the optical depth \\citep{1997ApJ...488....1Z}. Measurements by the WMAP satellite of these effects indicate that the redshift of reionization is $z_r\\approx11\\pm4$ \\citep{2007ApJS..170..377S}, assuming an instantaneous transition. Lyman-$\\alpha$ absorption by neutral hydrogen is visible in the spectra of many high-redshift quasars and, thus, offers the second currently feasible probe of the ionization history of the IGM. Continuum emission from quasars is redshifted as it travels through the expanding Universe to the observer. Neutral hydrogen along the line of sight creates absorption features in the continuum at wavelengths corresponding to the local rest-frame wavelength of the Lyman-$\\alpha$ line. Whereas CMB measurements place an integrated constraint on reionization, quasar absorption line studies are capable of probing the ionization history in detail along the sight-lines. There is a significant limitation to this approach, however. The Lyman-$\\alpha$ absorption saturates at very low fractions of neutral hydrogen (of order $x_{HI} \\approx 10^{-4}$). Nevertheless, results from these studies have been quite successful and show that, while the IGM is highly ionized below $z\\lesssim6$ (with typical $x_{HI}\\lesssim10^{-5}$), a significant amount of neutral hydrogen is present above, although precisely how much remains unclear \\citep{2001ApJ...560L...5D, 2001AJ....122.2850B, 2002AJ....123.1247F, 2003AJ....125.1649F, 2004Natur.427..815W, 2006AJ....132..117F}. The existing CMB and quasar absorption measurements are somewhat contradictory. Prior to these studies, the reionization epoch was assumed generally to be quite brief, with the transition from an IGM filled with fully neutral hydrogen to an IGM filled with highly ionized hydrogen occurring very rapidly. These results, however, open the possibility that the ionization history of the IGM may be more complicated than previously believed \\citep{2003ApJ...595....1H, 2003ApJ...591...12C, 2003MNRAS.344..607S, 2004ApJ...604..484M}. Direct observations of the 21~cm (1420~MHz) hyperfine transition line of neutral hydrogen in the IGM during the reionization epoch would resolve the existing uncertainties and reveal the evolving properties of the IGM. The redshifted 21~cm signal should appear as a faint, diffuse background in radio frequencies below $\\nu<200$~MHz for redshifts above $z>6$ (according to $\\nu=1420/[1+z]$~MHz). For diffuse gas in the high-redshift ($z\\approx10$) IGM, the expected unpolarized differential brightness temperature of the redshifted 21~cm line relative to the pervasive CMB is readily calculable from basic principles and is given by \\citep[their \\S~2]{2004ApJ...608..622Z} \\begin{equation} \\begin{array}{rl} \\label{eqn_intro_temp} \\delta T_{21}(\\vec{\\theta}, z) \\approx~& 23~(1+\\delta)~x_{HI} \\left ( 1 - \\frac{T_\\gamma}{T_S} \\right ) \\\\ & \\times \\left ( \\frac{\\Omega_b~h^2}{0.02} \\right ) \\left [ \\left ( \\frac{0.15}{\\Omega_m~h^2} \\right ) \\left ( \\frac{1+z}{10} \\right ) \\right ]^{1/2} \\mbox{mK}, \\end{array} \\end{equation} where $\\delta(\\vec{\\theta},z)$ is the local matter over-density, $x_{HI}(\\vec{\\theta},z)$ is the neutral fraction of hydrogen in the IGM, $T_\\gamma(z) = 2.73~(1+z)$~K is the temperature of CMB at the redshift of interest, $T_S(\\vec{\\theta},z)$ is the spin temperature that describes the relative population of the ground and excited states of the hyperfine transition, and $\\Omega_b$ is the baryon density relative to the critical density, $\\Omega_m$ is the total matter density, and $h$ specifies the Hubble constant according to $H_0=100~ h$~km~s$^{-1}$~Mpc$^{-1}$. From Equation~\\ref{eqn_intro_temp}, we see that perturbations in the local density, spin temperature, and neutral fraction of hydrogen in the IGM would all be revealed as fluctuations in the brightness temperature of the observed redshifted 21~cm line. The differential brightness temperature of the redshifted 21~cm line is very sensitive to the \\hi spin temperature. When the spin temperature is greater than the CMB temperature, the line is visible in emission. For $T_S \\gg T_\\gamma$, the magnitude of the emission saturates to a maximum (redshift-dependent) brightness temperature that is about 25 to 35~mK for a mean-density, fully neutral IGM between redshifts 6 and 15, assuming a $\\Lambda$CDM cosmology with $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $\\Omega_b=0.04$, and $h=0.7$. At the other extreme, when the spin temperature is very small and $T_S \\ll T_\\gamma$, the line is visible in absorption against the CMB with a potentially very large (and negative) relative brightness temperature. A number of factors are involved in predicting the typical differential brightness temperature of the redshifted 21~cm line as a function of redshift. In particular, the spin temperature must be treated in detail, including collisional coupling between the spin and kinetic temperatures of the gas, absorption of CMB photons, and heating by ultra-violet radiation from the first luminous sources. We direct the reader to \\citet{2006PhR...433..181F} for a good introduction to the topic. The results of several efforts to predict the evolution of the differential brightness temperature of the redshifted 21~cm line have yielded predictions that are generally consistent in overall behavior, but vary highly in specific details \\citep{1997ApJ...475..429M, 1999A&A...345..380S, 2004ApJ...608..611G, 2006MNRAS.371..867F}. These models tend to agree that, for a finite period at sufficiently high redshifts ($z\\gtrsim20$), the \\hi hyperfine line should be seen in absorption against the CMB, with relative brightness temperatures of up to $|\\delta T_b|\\lesssim100$~mK. This is because the IGM initially cools more rapidly than the CMB following recombination \\citep{1994ApJ...427...25S, 1997ApJ...475..429M}. During this period, fluctuations in the differential brightness temperature of the redshifted 21~cm background should track the underlying baryonic matter density perturbations \\citep{1972A&A....20..189S, 1979MNRAS.188..791H, 1990MNRAS.247..510S, 2002ApJ...572L.123I, 2003MNRAS.341...81I, 2004PhRvL..92u1301L, 2005ApJ...626....1B}. Eventually, however, the models indicate that the radiation from the first generations of luminous sources will elevate the spin temperature of neutral hydrogen in the IGM above the CMB temperature and the redshifted 21~cm line should be detected in emission with relative brightness temperatures up to the expected maximum values (of order $25$~mK). Finally, during the reionization epoch, the neutral hydrogen becomes ionized, leaving little or no gas to produce the \\hi emission, and the apparent differential brightness temperature of the redshifted 21~cm line falls to zero as reionization progresses. As the gas is ionized, a unique pattern should be imprinted in the redshifted 21~cm signal that reflects the processes responsible for the ionizing photons and that evolves with redshift as reionization progresses \\citep{1997ApJ...475..429M, 2000ApJ...528..597T, 2003ApJ...596....1C, 2004ApJ...608..622Z, 2004ApJ...613...16F}. The details of the specific timing, duration, and magnitude of these features remains highly variable between theoretical models due largely to uncertainties about the properties of the first luminous sources. Measuring the brightness temperature of the redshifted 21~cm background could yield information about both the global and the local properties of the IGM. Determining the average brightness temperature over a large solid angle as a function of redshift would eliminate any dependence on local density and temperature perturbations and constrain the evolution of the product $\\overline{x_{HI}(1-T_\\gamma/T_S)}$, where we use the bar to denote a spatial average. During the reionization epoch, it is, in general, believed to be a good approximation to assume that $T_S\\gg T_\\gamma$ and, therefore, that the brightness temperature is proportional directly to $\\bar{x}_{HI}$. Global constraints on the brightness temperature of the redshifted 21~cm line during the EOR, therefore, would directly constrain the neutral fraction of hydrogen in the IGM. Such constraints would provide a basic foundation for understanding the astrophysics of reionization by setting bounds on the duration of the epoch, as well as identifying unique features in the ionization history (for example if reionization occurred in two phases or all at once). They would also yield improvements in estimates of the optical depth to CMB photons and, thus, would help to break existing degeneracies in CMB measurements between the optical depth and properties of the primordial matter density power spectrum \\citep{2006PhRvD..74l3507T}. \\begin{figure} \\centering \\includegraphics[width=20pc]{f1_color.eps} \\caption[Photograph of EDGES deployed at Mileura Station]{ \\label{f_edges_photos} EDGES deployed at Mileura Station in Western Australia. The left panel shows the full antenna and ground screen in the foreground and the analog-to-digital conversion and data acquisition module in the background. The right panel is a close-up view of the amplifier and switching module connected directly to the antenna (through the balun).} \\end{figure} For these reasons, several efforts are underway to make precise measurements of the radio spectrum below $\\nu<200$~MHz ($z>6$). In this paper, we report on the initial results of the Experiment to Detect the Global EOR Signature (EDGES). In \\S~\\ref{s_edges_method}, we describe the specific approach used for EDGES to address the issue of separating the redshifted 21 signal from the foreground emission. We then give an overview of the EDGES system in \\S~\\ref{s_edges_system}, followed by the results of the first observing campaign with the system in \\S~\\ref{s_edges_results}, along with a discussion of the implications for future single-antenna measurements. ", "conclusions": "In principle, useful measurements of the redshifted 21~cm background can be carried out with a small radio telescope. These measurements would be fundamental to understanding the evolution of the IGM and the EOR. In particular, the global evolution of the mean spin temperature and mean ionization fraction of neutral hydrogen in the high redshift IGM could be constrained by very compact instruments employing individual radio antennas. We have reported preliminary results to probe the reionization epoch based on this approach from the first observing campaign with the EDGES system. These observations were limited by systematic effects that were an order of magnitude larger than the anticipated signal and, thus, ruled out only an already unlikely range of parameter space for the differential amplitude of the redshifted 21~cm brightness temperature and for the duration of reionization. Nevertheless, the results of this experiment indicate the viability of the simple global spectrum approach. Building on the experiences of these initial efforts, modifications to the EDGES system are underway to reduce the residual systematic contribution in the measured spectrum and to expand the frequency coverage of the system down to 50~MHz or lower in order to place constraints on the anticipated transition of the hyperfine line from absorption to emission as the IGM warms before the EOR. Constraining the redshift and intensity of this feature would be very valuable for understanding the heating history of the IGM and, since the transition has the potential to produce a step-like feature in the redshifted 21~cm spectrum with a magnitude over 100~mK (up to a factor of 4 larger than the amplitude of the step during the reionization epoch), it may be easier to identify than the transition from reionization---although the sky noise temperature due to the Galactic synchrotron foreground increases significantly at the lower frequencies, as well. Through these and other global spectrum efforts, the first contribution to cosmic reionization science from measurements of the redshifted 21~cm background will hopefully be achieved in the near future." }, "0710/0710.1175_arXiv.txt": { "abstract": "{The physical mechanism responsible for the short outbursts in a recently recognized class of High Mass X--ray Binaries, the Supergiant Fast X--ray Transients (SFXTs), is still unknown. Two main hypotheses have been proposed to date: the sudden accretion by the compact object of small ejections originating in a clumpy wind from the supergiant donor, or outbursts produced at (or near) the periastron passage in wide and eccentric orbits, in order to explain the low ($\\sim$10$^{32}$~erg~s$^{-1}$) quiescent emission. Neither proposed mechanisms seem to explain the whole phenomenology of these sources. } {Here we propose a new explanation for the outburst mechanism, based on the X--ray observations of the unique SFXT known to display periodic outbursts, IGR~J11215-5952. } {We performed three Target of Opportunity observations with \\sw, \\xmm\\ and \\inte\\ at the time of the fifth outburst, expected on 2007 February 9. \\sw\\ observations of the February 2007 outburst have been reported elsewhere. Another ToO with \\sw\\ was performed in July 2007, in order to monitor the supposed ``apastron'' passage. } {\\xmm\\ observed the source on 2007 February 9, for 23~ks, at the peak of the outburst, while \\inte\\ started the observation two days later, failing to detect the source, which had already undergone the decaying phase of the fast outburst. The Swift campaign performed in July 2007 reveals a second outburst occurring on 2007 July 24, as bright as that observed about 165~days before. } {The new X--ray observations allow us to propose an alternative hypothesis for the outburst mechanism in SFXTs, linked to the possible presence of a second wind component, in the form of an equatorial disk from the supergiant donor. We discuss the applicability of the model to the short outburst durations of all other Supergiant Fast X--ray Transients, where a clear periodicity in the outbursts has not been found yet. The new outburst from \\src\\ observed in July suggests that the true orbital period is $\\sim$165~days, instead of 329~days, as previously thought. } ", "introduction": "The Galactic plane monitoring performed by the \\inte\\ satellite in the last 5 years has allowed the discovery of a number of new High Mass X--ray Binaries (HMXBs). Several of these new sources are intrinsically highly absorbed and were difficult to discover with previous missions (e.g. IGR~J16318--4848, \\citealt{Walter2003}). Others are transient HMXBs (associated with OB supergiant) displaying short outbursts (few hours, typically less than a day; \\citealt{Sguera2005}), and form the recently recognized new class of Supergiant Fast X--ray Transients (SFXTs). \\object{IGR~J11215--5952} is a hard X--ray transient discovered by \\inte\\ during a fast outburst in April 2005 \\citep{Lubinski2005}. The short duration of this outburst led \\citet{Negueruela2005a} to propose that \\src\\ could be a new member of the class of Supergiant Fast X-ray Transients. The optical counterpart is indeed a B-type supergiant, \\object{HD~306414} located at a distance of 6.2~kpc (\\citealt{Negueruela2005b}, \\citealt{Masetti2006}, \\citealt{Steeghs2006}). From the analysis of archival \\inte\\ observations and the discovery of two previously unnoticed outbursts, a recurrence period in the X--ray activity of $\\sim$330~days has been found \\citep*[hereafter Paper I]{SidoliPM2006}, likely linked to the orbital period of the binary system. This periodicity was later confirmed by the fourth outburst from \\src\\ observed with $RXTE/PCA$ on 2006 March 16--17, 329~days after the previous one \\citep{Smith2006a}. The \\inte\\ spectrum was well fitted by a hard power-law with a high energy cut-off around 15~keV (Paper~I). Assuming a distance of 6.2 kpc, the peak fluxes of the outbursts correspond to a luminosity of $\\sim 3 \\times$10$^{36}$~erg~s$^{-1}$ (5--100~keV; Paper~I). The $RXTE/PCA$ observations showed a possible pulse period of $\\sim$195\\,s \\citep{Smith2006b}, later confirmed during the February 2007 outburst, yielding P=$186.78\\pm0.3$\\,s \\citep{Swank2007}. All these findings confirmed \\src\\ as a member of the class of SFXTs, and the first one displaying periodic outbursts. Based on the known periodicity, an outburst was expected for 2007 February 9. This allowed us to obtain several Target of Opportunity (ToO) observations with $Swift/XRT$, $XMM-Newton$ and $INTEGRAL$. The $Swift/XRT$ results of the February 2007 outburst have been reported in \\citealt{Romano2007} (hereafter Paper~II). Here we report the results of the \\xmm\\ and \\inte\\ observations of the February 2007 outburst, and of a $Swift/XRT$ campaing performed in July 2007, in order to monitor the supposed apastron passage. ", "conclusions": "\\label{conclusions} We have proposed here a new explanation for the short outbursts in SFXTs, based on our results of a monitoring campain of \\src. An equatorial wind from the supergiant companion is suggested, based on the narrow and steep shape of the X--ray lightcurve observed during the latest outburst. The short outburst is suggestive of a deviation from coplanarity ($\\theta$$>$0) of the equatorial plane of the companion with the orbital plane (as, e.g., in PSR~B1259-63, \\citealt{Wex1998} or in PSR~J0045--7319, \\citealt{Kaspi1994}), and of a some degree of eccentricity (e$>$0). Both orbital eccentricity and no-coplanarity can be explained by a substantial supernova ``kick'' (e.g. \\citealt{Eggleton2001}) at birth. This could suggest that SFXTs are likely young systems, probably younger than persistent HMXBs. We are aware that the derived quantities for the proposed supergiant equatorial disk are highly uncertain and still speculative. This uncertainty clearly emerges from the fact that the observed X--ray lightcurve can be reproduced with both orbital periods (329~days and 164.5~days), different eccentricities and different wind properties. A determination of the orbital eccentricity of the system, for example, would be essential in better constraining the expected properties of the wind which match with the level of X--ray emission during the outbursts. These considerations highlight the need for an as complete as possible deep monitoring of the X--ray emission along the orbit, together with optical observations in order to constrain the supergiant wind properties. \\begin{figure*} \\vbox{ \\includegraphics[height=6.5cm,angle=0]{8137fig10a.ps} \\includegraphics[height=6.5cm,angle=0]{8137fig10b.ps} \\includegraphics[height=6.5cm,angle=0]{8137fig10c.ps} \\includegraphics[height=6.5cm,angle=0]{8137fig10d.ps} \\includegraphics[height=6.5cm,angle=0]{8137fig10e.ps} \\includegraphics[height=6.5cm,angle=0]{8137fig10f.ps}} \\caption[]{Results of the proposed model compared with the $Swift/XRT$ lightcurve ({\\em upper panels}) the expected luminosity variations ({\\em middle panels}) and wind density along the whole neutron star orbit ({\\em lower panels}) in two cases: {\\em on the left} geometry case ``a'' (with orbital period of 164.5~days and an assumed eccentricity of 0.4) is shown (see Fig.~\\ref{fig:geom}), while {\\em on the right} the results for case ``b'' are reported (orbital period of 329~days, circular orbit and two similar outbursts per orbit). Both models assume a blue supergiant with a mass of 39~M$_\\odot$ and radius of 42~R$_\\odot$, a ``polar wind'' component \\textbf{(``PW'')} with a terminal velocity of 1800~km~s$^{-1}$. The X--ray lightcurve observed with $Swift/XRT$ is better reproduced assuming a ``polar wind'' mass loss rate of 5$\\times$10$^{-6}$~M$_\\odot$~yrs$^{-1}$ for case ``a'' and 9$\\times$10$^{-7}$ ~M$_\\odot$~yrs$^{-1}$ for case ``b''. The second wind component, in form of an equatorial disk (``ED''), has a variable velocity ranging from 750~km~s$^{-1}$ to 1400~km~s$^{-1}$ (for case ``a''), and from 850~km~s$^{-1}$ to 1600~km~s$^{-1}$ for case ``b''. The wind density profiles assumed here are reported in the two lower panels. For a magnetic field of 10$^{12}$~G the centrifugal barrier is open along all the orbit in both cases. Dashed line in the first upper right figure shows the model prediction based on a ten times smaller wind density with respect to that assumed to reproduce the peak luminosity (maintaining fixed all the other parameters). } \\label{fig:model} \\end{figure*} \\onltab{2}{ % \\begin{table*} \\begin{center} \\caption{Observation log.} \\label{igr112:tab:alldata2} \\begin{tabular}{lllll} \\hline \\hline \\noalign{\\smallskip} Sequence & Start time (MJD) & Start time (UT) & End time (UT) & Net Exposure$^{\\mathrm{a}}$ \\\\ & & (yyyy-mm-dd hh:mm:ss) & (yyyy-mm-dd hh:mm:ss) &(s) \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 00030881024 & 54256.7449\t &\t 2007-06-05 17:52:35\t &\t 2007-06-05 21:15:58\t &\t 1519 \\\\ 00030881025 & 54263.4918\t &\t 2007-06-12 11:48:11\t &\t 2007-06-12 16:57:58\t &\t 1522 \\\\ 00030881026 & 54270.3808\t &\t 2007-06-19 09:08:17\t &\t 2007-06-19 11:05:57\t &\t 1843 \\\\ 00030881027 & 54277.1485\t &\t 2007-06-26 03:33:52\t &\t 2007-06-26 05:27:57\t &\t 2109 \\\\ 00030881028 & 54284.8571\t &\t 2007-07-03 20:34:14\t &\t 2007-07-03 23:59:57\t &\t 2347 \\\\ 00030969001 & 54294.5627\t &\t 2007-07-13 13:30:15\t &\t 2007-07-14 07:10:56\t &\t 2126 \\\\ 00030881030 & 54298.0425\t &\t 2007-07-17 01:01:12\t &\t 2007-07-17 02:58:57\t &\t 2069 \\\\ 00030881031 & 54305.0012 \t & 2007-07-24 00:01:42 & 2007-07-24 13:16:55 & 1551 \\\\ 00030881032 & 54312.2343 & 2007-07-31 05:37:23 & 2007-07-31 10:26:56 & 1695 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{center} \\begin{list}{}{} \\item[$^{\\mathrm{a}}$] The exposure time is spread over several snapshots (single continuous pointings at the target) during each observation. \\end{list} \\end{table*} } %" }, "0710/0710.2189.txt": { "abstract": "{Minimal walking technicolor models can provide a nontrivial solution for cosmological dark matter, if the lightest technibaryon is doubly charged. Technibaryon asymmetry generated in the early Universe is related to baryon asymmetry and it is possible to create excess of techniparticles with charge ($-2$). These excessive techniparticles are all captured by $^4He$, creating \\emph{techni-O-helium} $tOHe$ ``atoms'', as soon as $^4He$ is formed in Big Bang Nucleosynthesis. The interaction of techni-O-helium with nuclei opens new paths to the creation of heavy nuclei in Big Bang Nucleosynthesis. Due to the large mass of technibaryons, the $tOHe$ ``atomic'' gas decouples from the baryonic matter and plays the role of dark matter in large scale structure formation, while structures in small scales are suppressed. %Due to Nuclear interactions with matter slow down cosmic techni-O-helium in Earth below the threshold of underground dark matter detectors, thus escaping severe CDMS constraints. On the other hand, these nuclear interactions %of techni-O-helium are not sufficiently strong to exclude this form of Strongly Interactive Massive Particles by constraints from the XQC experiment. Experimental tests of this hypothesis are possible in search for $tOHe$ in balloon-borne experiments (or on the ground) and for its charged techniparticle constituents in cosmic rays and accelerators. The $tOHe$ ``atoms'' can cause cold nuclear transformations in matter and might form anomalous isotopes, offering possible ways to exclude (or prove?) their existence.} ", "introduction": "The question of the existence of new quarks and leptons is among the most important in the modern high energy physics. This question has an interesting cosmological aspect. If these quarks and/or charged leptons are stable, they should be present around us and the reason for their evanescent nature should be found. Recently, at least three elementary particle frames for heavy stable charged quarks and leptons were considered: (a) A heavy quark and heavy neutral lepton (neutrino with mass above half the $Z$-boson mass) of a fourth generation \\cite{N,Legonkov}, which can avoid experimental constraints \\cite{Q,Okun}, and form composite dark matter species \\cite{I,lom,KPS06,Khlopov:2006dk}; (b) A Glashow's ``Sinister'' heavy tera-quark $U$ and tera-electron $E$, which can form a tower of tera-hadronic and tera-atomic bound states with ``tera-helium atoms'' $(UUUEE)$ considered as dominant dark matter \\cite{Glashow,Fargion:2005xz}; (c) AC-leptons, based on the approach of almost-commutative geometry \\cite{5,book}, that can form evanescent AC-atoms, playing the role of dark matter \\cite{5,FKS,Khlopov:2006uv}. In all these recent models, the predicted stable charged particles escape experimental discovery, because they are hidden in elusive atoms, composing the dark matter of the modern Universe. It offers a new solution for the physical nature of the cosmological dark matter. Here we show that such a solution is possible in the framework of walking technicolor models \\cite{Sannino:2004qp,Hong:2004td,Dietrich:2005jn,Dietrich:2005wk,Gudnason:2006ug,Gudnason:2006yj} and can be realized without an {\\it ad hoc} assumption on charged particle excess, made in the approaches (a)-(c). This approach differs from the idea of dark matter composed of primordial bound systems of superheavy charged particles and antiparticles, proposed earlier to explain the origin of Ultra High Energy Cosmic Rays (UHECR) \\cite{UHECR}. To survive to the present time and to be simultaneously the source of UHECR, superheavy particles should satisfy a set of constraints, which in particular exclude the possibility that they possess gauge charges of the standard model. The particles considered here, participate in the Standard Model interactions and we show how the problems, related to various dark matter scenarios with composite atom-like systems, can find an elegant solution on the base of the minimal walking technicolor model. The approaches (b) and (c) try to escape the problems of free charged dark matter particles \\cite{Dimopoulos:1989hk} by hiding opposite-charged particles in atom-like bound systems, which interact weakly with baryonic matter. However, in the case of charge symmetry, when primordial abundances of particles and antiparticles are equal, annihilation in the early Universe suppresses their concentration. If this primordial abundance still permits these particles and antiparticles to be the dominant dark matter, the explosive nature of such dark matter is ruled out by constraints on the products of annihilation in the modern Universe \\cite{Q,FKS}. Even in the case of charge asymmetry with primordial particle excess, when there is no annihilation in the modern Universe, binding of positive and negative charge particles is never complete and positively charged heavy species should retain. Recombining with ordinary electrons, these heavy positive species give rise to cosmological abundance of anomalous isotopes, exceeding experimental upper limits. To satisfy these upper limits, the anomalous isotope abundance on Earth should be reduced, and the mechanisms for such a reduction are accompanied by effects of energy release which are strongly constrained, in particular, by the data from large volume detectors. These problems of composite dark matter models \\cite{Glashow,5} revealed in \\cite{Q,Fargion:2005xz,FKS,I}, can be avoided, if the excess of only $-2$ charge $A^{--}$ %($(\\bar U \\bar U \\bar u)$, $(\\bar U \\bar U \\bar U)$) % or neutral %$(\\bar U u)$ particles is generated in the early Universe. Here we show that in walking technicolor models, technilepton and technibaryon excess is related to baryon excess and the excess of $-2$ charged particles can appear naturally for a reasonable choice of model parameters. It distinguishes this case from other composite dark matter models, since in all the previous realizations, starting from \\cite{Glashow}, such an excess was put by hand to saturate the observed cold dark matter (CDM) density by composite dark matter. After it is formed in Big Bang Nucleosynthesis, $^4He$ screens the $A^{--}$ charged particles in composite $(^4He^{++}A^{--})$ ``atoms''. These neutral primordial nuclear interacting objects saturate the modern dark matter density and play the role of a nontrivial form of strongly interacting dark matter \\cite{Starkman,McGuire:2001qj}. The active influence of this type of dark matter on nuclear transformations seems to be incompatible with the expected dark matter properties. However, it turns out that the considered scenario is not easily ruled out \\cite{FKS,I} and challenges the experimental search for techni-O-helium and its charged techniparticle constituents. The structure of the present paper is as follows. Starting with a review of possible dark matter candidates offered by the minimal walking technicolor model, we reveal the possibility for the lightest techniparticle(s) to have electric charge $\\pm 2$ (Section II). % In the framework of the considered approach, the %excess of techniparticles relative to their antiparticles can be %generated in early Universe and related to baryon asymmetry. In Section III we show how the minimal technicolor model can provide substantial excess of techniparticles with electric charge $-2$. %can be generated in the form of antitechnibaryons %$(\\bar{U}\\bar{U})^{-2}$, of technileptons $(\\zeta^{-2}$ or of %their mixture. In Section IV we show how all these $-2$ charge particles can be captured by $^4He$, after its formation in the Standard Big Bang Nucleosynthesis (SBBN), % and neutral %$^4He^{+2}(\\bar{U}\\bar{U})^{-2}$, $^4He^{+2}\\zeta^{-2}$ making neutral techni-O-helium ``atoms\" that can account for the modern dark matter density. Techni-O-helium catalyzes a path for heavy element formation in SBBN, but we stipulate in Section IV a set of arguments, by which the considered scenario can avoid immediate contradiction with observations. Gas of heavy techni-O-helium ``atoms\" decouples from the plasma and radiation only at a temperature about few hundreds eV, so that small scale density fluctuations are suppressed and gravitational instability in this gas develops more close to warm dark matter, rather than to cold dark matter scenario (subsection A of Section V). We further discuss in Section V the possibility to detect charged techniparticle components of cosmic rays (subsection B), effects of techni-O-helium catalyzed processes in Earth (subsection C), and possibilities of direct searches for techni-O-helium (subsection D). The problems, signatures, and possible experimental tests of the techni-O-helium Universe are considered in Section VI. Details of our calculations are presented in the Appendices 1 and 2. % In spite of technilepton $(\\zeta^{-2}$ excess, in the %difference with the case of technibaryons, the frozen out %abundance of antiparticles $(\\bar \\zeta^{+2}$ is not negligible. %Since $(\\bar \\zeta^{+2}$ represents a form of anomalous helium, it %might cause the problem of anomalous helium overproduction. %Cosmological evolution of technileptons and mechanisms of %suppression for primordial $(\\bar \\zeta^{+2}$ abundance are %considered in Appendix 2. ", "conclusions": "Discussion} In this paper we explored the cosmological implications of a walking technicolor model with doubly charged technibaryons and technileptons. The considered model escapes most of the drastic problems of the Sinister Universe \\cite{Glashow}, related to the primordial $^4He$ cage for $-1$ charge particles and a consequent overproduction of anomalous hydrogen \\cite{Fargion:2005xz}. These charged $^4He$ cages pose a serious problem for composite dark matter models with single charged particles, since their Coulomb barrier prevents successful recombination of positively and negatively charged particles. The doubly charged $A^{--}$ techniparticles considered in this paper, bind with $^4He$ in the techni-O-helium neutral states. %catalyzers of $AC$ binding and AC-leptons may thus escape this trap. To avoid overproduction of anomalous isotopes, an excess of $-2$ charged techniparticles over their antiparticles should be generated in the Universe. In all the previous realizations of composite dark matter scenarios, this excess was put by hand to saturate the observed dark matter density. In our paradigm, this abundance of techibaryons and/or technileptons is connected naturally to the baryon relic density. A challenging problem that we leave for future work is the nuclear transformations, catalyzed by techni-O-helium. The question about their consistency with observations remains open, since special nuclear physics analysis is needed to reveal what are the actual techni-O-helium effects in SBBN and in terrestrial matter. Another aspect of the considered approach is more clear. For reasonable values of the techiparticle mass, the amount of primordial $^4He$, bound in this atom like state is significant and should be taken into account in comparison to observations. The destruction of techni-O-helium by cosmic rays in the Galaxy releases free charged techniparticles, which can be accelerated and contribute to the flux of cosmic rays. In this context, the search for techniparticles at accelerators and in cosmic rays acquires the meaning of a crucial test for the existence of the basic components of the composite dark matter. At accelerators, techniparticles would look like stable doubly charged heavy leptons, while in cosmic rays, they represent a heavy $-2$ charge component with anomalously low ratio of electric charge to mass. To conclude, walking technicolor cosmology can naturally resolve most of problems of composite dark matter. Therefore, the model considered in this paper with stable $-2$ charged particles might provide a realistic physical basis for a composite dark matter scenario." }, "0710/0710.5560_arXiv.txt": { "abstract": "We describe the time- and position-dependent point spread function (PSF) variation of the Wide Field Channel (WFC) of the Advanced Camera for Surveys (ACS) with the principal component analysis (PCA) technique. The time-dependent change is caused by the temporal variation of the $HST$ focus whereas the position-dependent PSF variation in ACS/WFC at a given focus is mainly the result of changes in aberrations and charge diffusion across the detector, which appear as position-dependent changes in elongation of the astigmatic core and blurring of the PSF, respectively. Using $>400$ archival images of star cluster fields, we construct a ACS PSF library covering diverse environments of the $HST$ observations (e.g., focus values). We find that interpolation of a small number ($\\sim20$) of principal components or ``eigen-PSFs'' per exposure can robustly reproduce the observed variation of the ellipticity and size of the PSF. Our primary interest in this investigation is the application of this PSF library to precision weak-lensing analyses, where accurate knowledge of the instrument's PSF is crucial. However, the high-fidelity of the model judged from the nice agreement with observed PSFs suggests that the model is potentially also useful in other applications such as crowded field stellar photometry, galaxy profile fitting, AGN studies, etc., which similarly demand a fair knowledge of the PSFs at objects' locations. Our PSF models, applicable to any WFC image rectified with the Lanczos3 kernel, are publicly available. ", "introduction": "} Even in the absence of atmospheric turbulence, the finite aperture of Hubble Space Telescope ($HST$) causes light from a point source to spread at the focal plane with the diffraction pattern mainly reflecting the telescope's aperture and optical path difference function. Although the point-spread-function (PSF) of $HST$ is already far smaller than what one can achieve with any of the current ground-based facilities, astronomers' endless efforts to push to the limits of their scientific observations with $HST$ ever increase the demand for the better knowledge of the instrument's PSF. Especially, since the installation of the Advanced Camera for Surveys (ACS) on $HST$, there have been concentrated efforts to carefully monitor and understand the instrument's PSFs, and to utilize the unparalleled resolution and sensitivity of ACS in gravitational weak-lensing (e.g., Jee et al. 2005a; Heymans et al. 2005; Schrabback et al. 2007; Rhodes et al. 2007). Modeling the PSFs of ACS has proven to be non-trivial because of its complicated time- and position-dependent variation. The time-dependent change occurs due to the variation in the $HST$ focus, which relates to the constant shrinking of the secondary mirror truss structure and the thermal breathing of $HST$. The former is the main cause of the long-term focus change, and the secondary mirror position has been occasionally adjusted to compensate for this shrinkage (Hershey 1997). The latter is responsible for the short-term variation of the $HST$ focus and is affected by the instrument's earth heating, sun angle, prior pointing history, roll angle, etc. Even at a fixed focus value of $HST$, the PSFs of ACS also significantly change across the detector from the variation of the CCD thickness and the focal plane errors, which appear as position-dependent changes in charge diffusion and elongation of the astigmatic cores, respectively. The strategies to model these PSF variations can be categorized into two types: an empirical approach based on real stellar field observations and a theoretical prediction based on the understanding of the instrument's optics. The first method treats the optical system of the instrument nearly as a blackbox and mainly draws information from observed stellar images. Although the PSF variation pattern can be most straightforwardly described by the variation of the pixel intensity as a function of position (e.g., Anderson \\& King 2006), frequently orthogonal expansion of the observed PSFs (e.g., Lauer 2002; Bernstein \\& Jarvis 2002; Refregier 2003) have been utilized to make the description compact and tractable. On the other hand, the second approach mainly relies on the careful analysis of the optical configurations of the instrument and receives feedbacks from observations to fine-tune the existing optics model. The TinyTim software (Krist 2001) is the unique package of this type applicable to most instruments of $HST$. In this paper, we extend our previous efforts of the first kind (Jee et al. 2005a; 2005b; 2006; 2007) to describe the time- and position-dependent PSF variations of ACS/WFC now with the principal component analysis (PCA). In our previous work, we used ``shapelets'' (Bernstein \\& Jarvis 2002; Refregier 2003) to perform orthogonal expansion of the PSFs. Shapelets are the polar eigenfunctions of two-dimensional quantum harmonic oscillators, which form a highly localized orthogonal set. Although the decomposition of the stars with shapelets is relatively efficient and has proven to meet the desired accuracy for cluster weak-lensing analyses, the scheme is less than ideal in some cases. One important shortcoming is that it is too localized to capture the extended features of PSFs (Jee et al. 2007; also see \\textsection\\ref{section_basis_function}). In principle, the orthonormal nature of shapelets should allow us to represent virtually all the features of the target image when the number of basis functions are sufficiently large. However, this is not a viable solution not only because the convergence is slow, but also because the orthonormality breaks down in pixelated images for high orders as the function becomes highly oscillatory within a pixel. The PCA technique provides us with a powerful scheme to obtain the optimal set of basis functions from the data themselves. Unlike ``shapelets'', the basis functions derived from the PCA are by nature non-parametric, discrete, and highly customized for the given dataset. Therefore, it is possible to summarize the multi-variate statistics, with a significantly small number of basis functions (i.e., much smaller than the dimension of the problem). For example, PCA has been applied to the classification of object spectra in large area surveys (Connolly et al. 1995; Bromley et al. 1998; Madgwick et al. 2003). It has been shown that only a small number ($10\\sim 20$) of the basis functions or $eigenspectra$ are needed to reconstruct the sample. The application of PCA to the PSF decomposition is used by the Sloan Digital Sky Survey (SDSS) to model the PSF variations (e.g., Lupton et al. 2001; Lauer 2002). Jarvis \\& Jain (2004) used the PCA technique to describe the variation in the PSF pattern in the CTIO 75 square-degree survey for cosmic shear analyses. They fit the ``rounding'' kernel component with PCA, not the PSF shape directly. This scheme is motivated by their shear measurement technique (i.e., reconvolution to remove systematic PSF anisotropy). However, in the current study we choose to fit the PSF shapes directly because this is more general in the sense that the rounding kernel components are not uniquely determined for a given PSF. In addition, our PSF library generating the PSF shapes directly has more uses in other studies. We aim to construct a high-quality PSF library for the broadband ACS filters (F435W, F475W, F555W, F606W, F625W, F775W, F814W, and F850LP) from $>400$ archival stellar images, which sample a wide range of the $HST$ environments (e.g., the focus values). Our PSF models describe ACS PSFs in rectified images, specifically, drizzled using the Lanczos3 kernel with an output pixel scale of 0.05\\arcsec (see \\textsection\\ref{section_drizzling_kernel} for the justification of this choice). The results from this work are made publicly available on-line via the ACS team web site\\footnote{The full PSF library of ACS will become available at http://acs.pha.jhu.edu/$\\sim$mkjee/acs\\_psf/.}. We will present our works as follows. The justification and the basic mathematical formalism of PCA are briefed in \\textsection\\ref{section_PCA}. In \\textsection\\ref{section_application_acs}, we demonstrate how the technique can be applied to ACS data with some test results. Focus dependency of the ACS PSFs, comparison with TinyTim, and strategies to find matching templates are discussed in \\textsection\\ref{section_discussion} before we conclude in \\textsection\\ref{section_conclusion}. ", "conclusions": "} We showed that the time- and position-dependent ACS/WFC PSF can be robustly described through PCA. The PCA technique allows us to perform orthogonal expansion of the observed PSFs with as few as 20 eigen-PSFs derived from the data themselves. This method is superior to our previous shapelet-based decomposition of the PSFs, capturing more details of the diffraction pattern of the instrument PSF. By interpolating the position-dependent variation of the eigen-PSFs with 5th order polynomials, we are able to recover the observed pattern of the PSF ellipticity and width variation. Although the TinyTim software provides a good approximation of the observed PSFs, we demonstrate that there are some important mismatches between the TinyTim prediction and the real PSFs, which cannot be attributed to CTE degradation of WFC over time. The CTE charge trailing effect should be negligible for these bright high S/N stars, and we do not observe any long-term variation of the pattern (i.e., increasing elongation in parallel read-out direction with time) due to the CTE degradation. Because typical science observations require integration of one or more orbits in broadband filters, the background levels are high ($\\sim200$ $e^{-}$ for integration of one orbit). These high background photons are supposed to fill the charge traps and thus mitigate the CTE effects. Therefore, we argue that the CTE-induced elongation is not likely to limit the application of our PSF models extracted from short-exposure observations to long-exposure science images. We have compiled WFC PSFs from $>400$ stellar field observations, which span a wide range of $HST$ focus values. Although the current paper mainly deals with the ACS/WFC PSF issue in the context of weak-lensing analysis, we believe that our PSF model can be used in a wide range of the astronomical data analyses where the knowledge of the position-dependent WFC PSF is needed (e.g., crowded field stellar photometry, robust profile fitting of small objects, weak-lensing analyses, etc.). ACS was developed under NASA contract NAS5-32865, and this research was supported by NASA grant NAG5-7697." }, "0710/0710.0345_arXiv.txt": { "abstract": "We present results of a population synthesis study aimed at examining the role of spin-kick alignment in producing a correlation between the spin period of the first-born neutron star and the orbital eccentricity of observed double neutron star binaries in the Galactic disk. We find spin-kick alignment to be compatible with the observed correlation, but not to alleviate the requirements for low kick velocities suggested in previous population synthesis studies. Our results furthermore suggest low- and high-eccentricity systems may form through two distinct formation channels distinguished by the presence or absence of a stable mass transfer phase before the formation of the second neutron star. The presence of highly eccentric systems in the observed sample of double neutron stars may furthermore support the notion that neutron stars accrete matter when moving through the envelope of a giant companion. ", "introduction": "Recent observations of single and binary pulsars have sparked new questions and challenged accepted ideas on the formation of neutron stars (NSs) and the nature of supernova (SN) kicks. In particular, measurements of pulsar radio emission polarization and pulsar wind nebulae symmetry axis directions have provided increasingly compelling evidence for the alignment of pulsar proper motions and pulsar rotation axes \\citep[e.g.][]{2005MNRAS.364.1397J, 2006ApJ...639.1007W, 2006ApJ...644..445N, 2007ApJ...660.1357N, 2007ApJ...664..443R, 2007arXiv0708.4251J}. Assuming the proper motion and pulsar spin axis directions are representative of the natal kick and NS progenitor rotation axis, the alignment suggests that natal kicks are preferentially aligned with the progenitor's rotation axis. Moreover, the increasing sample of double neutron star (DNS) binaries has revealed a possible correlation between the spin period $P_{\\rm spin}$ of the first born NS and the binary orbital eccentricity $e$ (see Fig.~\\ref{f1}) \\citep{2005ASPC..328...43M, 2005ApJ...618L.119F}. Such a relation arises naturally for symmetric SN explosions, but is highly constraining for asymmetric explosions. Our aim in this paper is to examine the role of spin-kick alignment in establishing the observed $P_{\\rm spin}$--$e$ correlation. \\begin{figure} \\includegraphics[height=.24\\textheight]{observedpspinns1e.ps} \\caption{Orbital eccentricities and spin periods of observed DNSs in the Galactic disk. The dashed line represents the best-fitting log-linear curve $e=-1.7 + 1.2 \\log P_{\\rm spin}$.} \\label{f1} \\end{figure} ", "conclusions": "We investigated the role of spin-kick alignment in establishing a correlation between the spin period of the first-born NS and the orbital eccentricity of DNSs using the StarTrack binary population synthesis code and a simple prescription for the spin-up of a NS due to mass accretion. For DNSs forming through the standard evolutionary channel, spin-kick alignment is compatible with the observed $P_{\\rm spin}$--$e$ relation, but does not alleviate the requirement for low kick velocities proposed in previous population synthesis studies based on isotropic kick distributions. This is puzzling considering the stringent lower limits on the kick velocities imparted to the second-born NS in PSR\\,B1534+12 and PSR\\,B1913+16. Moreover, if low kicks are imparted to the second-born NS in DNSs, the standard formation channel cannot explain the formation of the high-eccentricity systems PSR\\,B1913+16 and PSR\\,J1811-1736. A possible resolution would be a dichotomous formation channel where low-eccentricty DNSs are formed through the standard formation channel with low kicks imparted to the second-born NS, and high-eccentricity systems are formed through a formation channel where no mass transfer occurs after the common envelope phase of the second NS's progenitor and \"normal\" kicks typical of isolated radio pulsars are imparted to the second-born NS. The low-eccentricity systems on the $P_{\\rm spin}$--$e$ relation can be formed with either polar or isotropic kicks. However, to obtain a $P_{\\rm spin}$--$e$ relation at high eccentricities, some spin-kick alignment is required to counteract the increased spread in the post-SN orbital eccentricities introduced by the larger kicks. We will explore this in the continuation of this investigation. \\begin{theacknowledgments} This work is partially supported by a Packard Foundation Fellowship, a NASA BEFS grant (NNG06GH87G), and a NSF CAREER grant (AST-0449558) to VK. \\end{theacknowledgments}" }, "0710/0710.3085_arXiv.txt": { "abstract": "We present the results of recent observations of phase-dependent variations in brightness designed to characterize the atmospheres of hot Jupiters. In particular, we focus on recent observations of the transiting planet HD~189733b at 8~\\micron~using the \\emph{Spitzer Space Telescope}, which allow us to determine the efficiency of the day-night circulation on this planet and estimate the longitudinal positions of hot and cold regions in the atmosphere. We discuss the implications of these observations in the context of two other successful detections of more sparsely-sampled phase variations for the non-transiting systems $\\upsilon$~And~b and HD~179949b, which imply a potential diversity in the properties of the atmospheres of hot Jupiters. Lastly, we highlight several upcoming \\emph{Spitzer} observations that will extend this sample to additional wavelengths and more transiting systems in the near future. ", "introduction": "There are currently more than 20 known transiting planetary systems, of which the majority are gas-giant planets orbiting extremely close ($<$0.05~A.U.) to their parent stars\\footnote{See http://vo.obspm.fr/exoplanetes/encyclo/encycl.html for the latest count}. These planets, known as ``hot Jupiters'', receive $>$10,000 times more radiation from their stars than Jupiter does from the sun, heating them to temperatures as high as 2000~K \\citep{har07}. The picture is further complicated by the fact that most of these planets are expected to be tidally locked, with permanent day and night sides. As a result, the equilibrium compositions and circulation patterns in these atmospheres are expected to differ significantly from those of the gas-giant planets in the solar system. It is not surprising, then, that simple models for the atmospheres of these planets \\citep[see, for example, ][]{seag05,bar05,fort06b,bur07b} can sometimes differ significantly in both their assumptions and the conclusions that they reach; this is particularly true of the circulation models discussed in \\S\\ref{models} Fortunately, it is possible to test the predictions of these models using observations of transiting systems. From the wavelength-dependent depth of the transit, when the planet passes in front of the star, we can search for features in the transmission spectrum of the planet near the day-night terminator \\citep{char02,vid03,vid04,knut07a,bar07,tin07,ehren07}. Similarly, by measuring the depth of the secondary eclipse, when the planet moves behind the star, we can characterize the properties of the emission from the day side of the planet \\citep{char05,dem05,dem06,dem07,demory07,grill07,rich07,har07,knut07b,knut07c}. Lastly, by measuring the changes in brightness over a significant fraction of a planet's orbit we can directly constrain the longitudinal temperature distribution across the planet's atmosphere \\citep{har06,cow07,knut07b}. This last type of observation is particularly informative for hot Jupiters, which may have extreme changes in temperature between the day and night sides of the planet. It is also a difficult observation to make in practice as it requires a very high precision over time scales of several days; this is why all of the successful detections to date have been made from space. Although the shape of the ingress and egress during secondary eclipse also contain information about the brightness distribution on the day side of the planet \\citep{will06,rau07}, these variations are smaller and take place on shorter time scales, and there have not been any clear detections of this effect. In this paper we focus on observations of phase variations, discussing the predictions of the current generation of atmospheric circulation models, the results of the three successful measurements of phase variations to date, and upcoming plans for more such observations in the near future. ", "conclusions": "The above discussion highlights the significant advances that have been made in the year since \\citet{har06} reported the first detection of the phase variation of an extrasolar planet. However, the current sample is limited, and there is clearly much to be gained from observations of additional transiting systems, for which the planetary radius, orbital inclination, and dayside flux are all well-constrained. There are a number of exciting programs scheduled for the upcoming \\emph{Spitzer} observing cycle which will address this issue. First, we have a program which will observe HD 189733b continuously at 24~\\micron~over the same half orbit as \\citet{knut07b}, which should probe a different region of the atmosphere than the previous 8~\\micron~observations. As a part of this same program we will also observe HD~209458b over half an orbit at both 8 and 24~\\micron, which will allow us to directly compare the wavelength-dependent properties of the atmospheric circulation for these two planets. To date all of the observations we have described have examined planets with effectively circular orbits; however, there is another promising set of upcoming observations by G. Laughlin and G. Bakos that will observe two highly eccentric planets (HD 80606b and HAT-P-2b) continuously over $\\sim30$ hours centered on periastron passage. Although only one of these planets, HAT-P-2b, is transiting, both observations should detect the rapid increase in brightness as the planetary atmosphere is flash-heated during this passage, providing a direct measurement of the radiative time constant. Of the two methods described above, it is the more time-intensive, continuous observations of transiting systems that have provided the most detailed information about circulation in the atmospheres of hot Jupiters to date. However, the detections of phase variations for $\\upsilon$~And~b and HD~179949b highlight the advantages of more sparsely-sampled observations in facilitating a much wider-ranging survey of a large number of planetary systems. These two approaches are inherently complementary, and we expect that both will make valuable contributions to our understanding of these unusual planets in the near future." }, "0710/0710.4170_arXiv.txt": { "abstract": "{We present the analysis and results of recent high-energy gamma-ray observations of the high energy-peaked BL Lac (HBL) object 1ES 1218+304 with the Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE). 1ES 1218+304 is an X-ray bright HBL at a redshift z=0.182. It has been predicted to be a gamma-ray emitter above 100 GeV, detectable by ground-based Cherenkov telescopes. Recently, this source has been detected by MAGIC and VERITAS, confirming these predictions. STACEE's sensitivity to astrophysical sources at energies above 100 GeV allows it to explore high energy sources such as X-ray bright active galaxies and gamma-ray bursts. We present results from STACEE observations of 1ES 1218+304 in the 2006 and 2007 observing seasons. } \\begin{document} ", "introduction": "Active galaxies of the ``blazar'' class include BL Lac objects and flat-spectrum radio quasars (FSRQs), and are characterized by non-thermal continuum emission that extends from radio to high energy gamma rays. The spectral energy distributions (SEDs) of these sources typically have two broad peaks, one at low energies (radio to X-ray) and the other at higher energies (keV to TeV). In the framework of relativistic jet models, these objects are highly beamed sources, emitting plasma in relativistic motion (e.g. \\cite{Urry&Padovani1995}). In blazar SEDs, the low energy peak is explained as synchrotron emission from high energy electrons in the jet, while the high energy peak is probably due to inverse Compton emission. Several competing ``leptonic'' and ``hadronic'' jet model explanations exist for the high energy emission (e.g. see \\cite{Boettcher2002} \\& \\cite{Muckeetal2003} for reviews), and further broadband observation of blazars are needed to distinguish between these models. Since the discovery of blazars as high energy gamma-ray sources by the Energetic Gamma-Ray Experiment Telescope (EGRET) on board the {\\sl Compton Gamma Ray Observatory} (CGRO) \\cite{Hartman1999} and the first detection of a TeV ($10^{12}$ eV) blazar by the Whipple Observatory (Mrk 421 \\cite{Punch1992}), the search has been on for more TeV blazars. The number count of TeV blazars is growing, with the advent of new generation atmospheric Cherenkov telescopes (ACTs) \\cite{Hessblazar19xx}. To date, almost all confirmed blazars detected at TeV energies are high-frequency-peaked BL Lac objects (HBLs), as opposed to quasars that constitute the majority of the EGRET detections. HBLs are a sub-class of BL Lac objects, characterized by lower luminosity than FSRQs and a synchrotron peak in the X-ray band \\cite{Urry&Padovani1995}. Blazars are categorized into different sub-classes based on the peak frequencies and the relative power in the low and high energy peaks of their SEDs \\cite{Fossati1998}. Given the high synchrotron peak frequencies of HBLs, indicating the presence of high energy electrons, these sources have been predicted to be good candidates for TeV emission, based on synchrotron self-Compton (SSC) emission models \\cite{Costamante&Ghisellini2002} as well as hadronic models \\cite{Mannheim1993}. Several of the ``extreme'' synchrotron BL Lacs \\cite{Costamante2001} have been detected at TeV energies, confirming these predictions. 1ES 1218+304 is an X-ray bright (flux at 1 keV $> 2$ $\\mu$Jy) HBL, categorized as an ``extreme'' BL Lac, and predicted to be a TeV source. The source was recently detected by both MAGIC \\cite{Albert2006} and VERITAS \\cite{Fortin2007}, at energies $>100$ GeV. At a redshift of $z=0.182$, 1ES 1218+304 is one of the most distant blazars detected to date. The source was never detected by EGRET, indicating that ACTs are sensitive to a different population of gamma-ray blazars than EGRET. The Solar Tower Atmospheric Cherenkov Effect Experiment (STACEE) is a ground-based experiment that is sensitive to gamma rays above 100 GeV. STACEE observations of AGN are motivated by the need to understand particle acceleration and emission mechanisms in blazars, as well as their interaction with the extragalactic background light (EBL). Despite what is already known, a great deal remains to be discovered regarding the physics of blazars. STACEE's extragalactic observing program has included both HBLs, as well as LBLs \\cite{Mukherjee2006, Mukherjee2005}. Recent observations of 1ES 1218+304 with STACEE were motivated by the detection of TeV emission from the source by MAGIC, providing further evidence that X-ray bright HBLs tend to be strong VHE sources. STACEE observations of 1ES 1218+304 were carried out in the 2006 and 2007 observing seasons. In this paper we present a summary of STACEE observations of the HBL 1ES 1218+304. ", "conclusions": "STACEE observed the high frequency-peaked BL Lac object 1ES 1218+304 in 2006 and 2007. After all cuts and padding 28.3 hr of data yielded an ON-source excess with a significance of $2.3\\sigma$ consistent with no detected flux. Simulated effective areas (Figure 3) were used to derive flux upper limits. For the combined 2006 and 2007 data sets the differential flux upper limit at the 99\\% confidence level was derived to be $< 5.2 \\times 10^{-6}$ m$^{-2}$ s$^{-1}$ TeV$^{-1}$ at 150 GeV, the energy threshold of STACEE. The upper limit was calculated assuming that the differential flux of photons follows a power law with an index of $-3.0$, as measured by MAGIC \\cite{Albert2006}. The STACEE upper-limit is shown overlaid on the MAGIC spectrum in Figure 4. These numbers are consistent with the spectrum measured by MAGIC \\cite{Albert2006}, and with the recent VERITAS results indicating a weak gamma-ray source at $\\sim5$\\% of the Crab flux \\cite{Fortin2007}. \\begin{figure} \\begin{center} \\includegraphics*[width=0.48\\textwidth,height=2.5in]{rmukherjee_0403_fig4.ps} \\end{center} \\caption{Gamma-ray spectrum of 1ES 1218+304, as measured by MAGIC (figure from \\cite{Albert2006}). The STACEE 99\\% flux upper limit is overlaid on the plot at 150 GeV, the energy threshold of STACEE, as obtained from detector simulations. The upper limit was calculated assuming the power-law spectral index of $-3.0$ measured by MAGIC. } \\label{fig4} \\end{figure} \\bigskip \\vskip 2in {\\bf Acknowledgements: } Many thanks go to the staff of the National Solar Tower Test Facility, who have made this work possible. This work was funded in part by the US National Science Foundation, the Natural Sciences and Engineering Research Council of Canada, Fonds Quebecois de la Recherche sur la Nature et les Technologies, the Research Corporation, and the University of California at Los Angeles. R. M. acknowledges support from NSF grant 0601112." }, "0710/0710.1080_arXiv.txt": { "abstract": "HCN and CO line diagnostics provide new insight into the OH megamaser (OHM) phenomenon, suggesting a dense gas trigger for OHMs. We identify three physical properties that differentiate OHM hosts from other starburst galaxies: (1) OHMs have the highest mean molecular gas densities among starburst galaxies; nearly all OHM hosts have $\\bar{n}({\\rm H}_2)=10^3$--$10^4$~cm$^{-3}$ (OH line-emitting clouds likely have $n({\\rm H}_2)>10^4$~cm$^{-3}$). (2) OHM hosts are a distinct population in the nonlinear part of the IR-CO relation. (3) OHM hosts have exceptionally high dense molecular gas fractions, $L_{\\rm HCN}/L_{\\rm CO}>0.07$, and comprise roughly half of this unusual population. OH absorbers and kilomasers generally follow the linear IR-CO relation and are uniformly distributed in dense gas fraction and $L_{\\rm HCN}$, demonstrating that OHMs are independent of OH abundance. The fraction of non-OHMs with high mean densities and high dense gas fractions constrains beaming to be a minor effect: OHM emission solid angle must exceed $2\\pi$ steradians. Contrary to conventional wisdom, IR luminosity does not dictate OHM formation; both star formation and OHM activity are consequences of tidal density enhancements accompanying galaxy interactions. The OHM fraction in starbursts is likely due to the fraction of mergers experiencing a temporal spike in tidally driven density enhancement. OHMs are thus signposts marking the most intense, compact, and unusual modes of star formation in the local universe. Future high redshift OHM surveys can now be interpreted in a star formation and galaxy evolution context, indicating both the merging rate of galaxies and the burst contribution to star formation. ", "introduction": "OH megamasers (OHMs) are rare luminous 18~cm masers associated with major galaxy merger-induced starbursts. The hosts of OHMs are (ultra)luminous IR galaxies ([U]LIRGs), and the OHM fraction in (U)LIRGs peaks at about 1/3 % in the highest luminosity mergers \\citep{darling02a}. It is not known whether all major mergers experience an OHM stage or what detailed physical conditions produce OHMs, but it is clear that OHMs are a radically different phenomenon from the aggregate OH maser emission associated with ``normal'' (Galactic) modes of star formation in galaxies. \\citet{lo05} posed a key question: why do $80\\%$ of LIRGs show no OHM activity? To reframe the question: given two merging systems with similar global IR and radio continuum properties in the same morphological stage of merging, why does one show OHM emission while the other does not? What is the difference between the two systems? Perhaps there is no difference and the fraction of OHMs among mergers simply reflects beaming or OH abundance. Or perhaps OHM activity depends on small-scale conditions that are decoupled from global properties of mergers. The provenance of OHM emission vis-\\`{a}-vis the host galaxy has been extensively investigated in the radio through X-ray bands by comparing samples of OHM galaxies to similarly selected non-masing control samples. For example, \\citet{darling02a} and \\citet{baan06} studied radio and IR properties vis-\\`{a}-vis the AGN versus starburst contributions to OHM activity, \\citet{baan1992} and \\citet{darling02a} investigated the OHM fraction in (U)LIRGs versus star formation rate and IR color, \\citet{baan1998} and \\citet{darling06} used optical spectral classification to distinguish populations and to quantify AGN fraction in OHM hosts, and \\citet{vignali05} conducted an X-ray study of the contribution of AGNs to OHM hosts. While some of these studies pointed to minor differences in statistical samples of OHM hosts versus nonmasing systems, they could not identify on a case-by-case basis which systems would harbor OHMs and which would not based on any observable quantity except the OH line itself. Theoretical modeling of OHM formation has seen a recent renaissance: \\citet{parra2005} model the $\\sim50$~pc molecular torus in III~Zw~35 and show how OHM emission is a stochastic amplification of unsaturated emission by multiple overlapping clouds, and \\citet{lockett07} show how the general excitation of OHMs is fundamentally different from Galactic OH maser emission and predict that a single excitation temperature governs all 18~cm OH lines. While the physics of OHMs is crystallizing, and models predict that beaming is not likely to be the dominant factor in the OHM fraction among (U)LIRGs, it remains unclear on a case-by-case basis what conditions found in starbursts drive or prohibit OHM formation. Here we describe a dense gas trigger for OHM formation, at last identifying physical observable properties that differentiate OHMs from nonmasing mergers. We identify OHMs, OH absorbers, OH kilomasers, and OH non-detections in the Gao \\& Solomon (2004a; hereafter GS04a) sample (\\S \\ref{sec:sample}) and employ CO($1-0$) and HCN($1-0$) molecular gas tracers to show that while OH absorbers appear nearly uniformly distributed in $L_{\\rm IR}$ and $L_{\\rm HCN}$, OHMs represent the {\\it majority} of the nonlinear population in the IR-CO relation (\\S \\ref{sec:results}). In combination with a Kennicutt-Schmidt-based star formation model of CO line emission by \\citet{krumholz07}, we identify a high mean molecular density driving OHM emission, and from the HCN/CO ratio we find that OHM galaxies are exclusively high dense gas fraction starbursts (\\S \\ref{sec:results}). Now that we can at last observe quantities that are highly predictive of OHM activity, we can employ OHMs at high redshifts as probes of major galaxy mergers and extreme star formation (\\S \\ref{sec:discussion}). ", "conclusions": "We have identified three closely related physical properties that differentiate OHMs from other starburst galaxies: OHM hosts have the highest mean molecular gas densities, they are a distinct population in the nonlinear part of the IR-CO relation, and they reside in galaxies with exceptionally high dense molecular gas fractions. We conclude that molecular gas must be concentrated and massive in order to reach the mean density required to form an OHM in a galactic nucleus. IR luminosity is not a condition for OHM formation; both star formation and OHM activity are consequences of the tidal density enhancements accompanying galaxy interactions. The fraction of OHMs in dense starbursts constrains OHM beaming to be a minor effect: OHM solid angle emission must be greater than $2\\pi$ steradians. These conclusions are in good agreement with the stochastic cloud-cloud overlap amplification model by \\citet{parra2005}. The rather uniform distribution of OH absorbers in IR, HCN, and CO luminosity suggests that OH abundance is not a significant factor in OHM formation. The main caveat to these conclusions is that the sample of OHMs with HCN observations remains small, and should be expanded, particularly to higher redshifts to include ``typical'' OHMs. OHMs are signposts of the most intense, compact, and unusual modes of star formation in the local universe, and surveys for OHMs will now provide detailed information about the detected host galaxies and their mode of star formation. The missing datum required for a complete interpretation of OHM surveys, however, is the OHM lifetime." }, "0710/0710.4941_arXiv.txt": { "abstract": "The bulk composition of an exoplanet is commonly inferred from its average density. For small planets, however, the average density is not unique within the range of compositions. Variations of a number of important planetary parameters---which are difficult or impossible to constrain from measurements alone---produce planets with the same average densities but widely varying bulk compositions. We find that adding a gas envelope equivalent to 0.1\\%-10\\% of the mass of a solid planet causes the radius to increase 5-60\\% above its gas-free value. A planet with a given mass and radius might have substantial water ice content (a so-called ocean planet) or alternatively a large rocky-iron core and some H and/or He. For example, a wide variety of compositions can explain the observed radius of GJ 436b, although all models require some H/He. We conclude that the identification of water worlds based on the mass-radius relationship alone is impossible unless a significant gas layer can be ruled out by other means. ", "introduction": "Out of over 250 exoplanets known to date, over 20 are known to transit their stars. Transiting planets are important because we can derive the precise mass and radius, and can begin to determine other planetary properties, such as the bulk composition. Much attention has been given to ``ocean planets'' or ``water worlds'', planets composed mostly of solid water \\citep{kuch2003, lege2004}. If a water world is found close to a star, it would be strong evidence for migration because insufficient volatiles exist near the star for {\\it in situ} formation. The proposed identification of water worlds is through transits. From a measured mass and radius a low-density water planet could potentially be identified. We examine the possibility that water worlds cannot be uniquely identified based on the mass and radius of a transiting planet. An alternative interpretation could be a rocky planet with a thick hydrogen-rich atmosphere. Most authors have assumed that solid planets in the 5 to 10 $M_{\\oplus}$ range have an insignificant amount of hydrogen \\citep{vale2006, vale2007a, fort2007, seag2007, sels2007, soti2007}. Exoplanets have, however, contradicted our basic assumptions before. Notable examples include: the existence of hot Jupiters; the predominance of giant planets in eccentric orbits; and the gas-rock hybrid planet HD~149026b with its $\\sim 60 M_{\\oplus}$ core and $\\sim 30 M_{\\oplus}$ H/He envelope \\citep{sato2005}. We adopt the idea that a wide range of atmospheric formation and loss mechanisms exist and can lead to a range of atmosphere masses on different exoplanets. We explore the mass-radius relationship for the lowest-mass exoplanets yet detected ($\\sim$~5-20 $M_{\\oplus}$) in order to identify potential ambiguities that result from the presence of a massive atmosphere. We explore atmospheres ranging from $\\sim 10^{-3} M_{\\oplus}$ (10 $\\times$ Venus' atmospheric mass) to $\\sim 1 M_{\\oplus}$ (the estimated mass of Uranus' and Neptune's H/He \\citep{guil2005, podo1995, hubb1991}), with a focus on the smaller mass range. We also explore potential compositions for the transiting Neptune-size planet GJ~436b \\citep{butl2004, gill2007a}. ", "conclusions": "Our fiducial planet consists of a 30\\% Fe core and a 70\\% MgSiO$_3$ mantle, roughly analogous to Earth. We used a H/He mixture with helium mass fraction $Y=0.28$ (the He mass fraction of the solar nebula). We chose $T_{eq}=300$~K, based on the observation that a planet around an M-dwarf at an orbital distance of 0.1 AU has a similar equilibrium temperature to Earth (assuming similar albedos). We set $T_{eff}=30$~K, similar to Earth and Uranus\\footnote{Earth has 44 $\\times 10^{12}$~ W \\citep{poll1993} and Uranus has $340 \\times 10^{12}$~W of energy flow \\citep{pear1990}.}. For the atmospheric parameters, we fixed $\\mu_0 = \\cos 60^{\\circ}$ and $\\gamma=0.1$ to represent radiation absorbed deep in the atmosphere\\footnote{In comparison $\\gamma=10$ would correspond to absorption high in the atmosphere.}. We later investigate variations on $Y$, $T_{eff}$, $T_{eq}$ and $\\gamma$. Figure~1 shows a plot of the mass-radius relationship for fiducial planets of masses 5, 10, 15, and 20 $M_{\\oplus}$. For each planet mass, we added atmospheres ranging in mass from 0.001-1 $M_{\\oplus}$. A robust finding for all models is that a small amount of gas creates a large radius increase. While this result is expected, the radius increase is far more dramatic than anticipated. For example, an H/He atmosphere of $\\sim 0.001$ by mass---only ten times greater than Venus' atmospheric mass fraction---is required for a noticeable radius increase. As seen in Figure~1, adding a hydrogen-helium atmosphere with just 0.1\\% of the mass of a 10 $M_{\\oplus}$ rocky planet results in a 5\\% increase in the planetary radius---within a measurement precision that has been obtained for currently known transiting planets. \\begin{figure} \\plotone{f1} \\caption{The increase in radius due to adding H/He to a solid planet. A H/He layer of 0.002-1 $M_{\\oplus}$ is added to a solid planet of 5, 10, 15, or 20 $M_{\\oplus}$, with fiducial model parameters (30\\% Fe and 70\\% MgSiO$_3$). The black points are for atmospheres at 0.01 $M_{\\oplus}$ and every 0.1 $M_{\\oplus}$ afterwards. The mass-radius relationship of solid planets with no gas is plotted for comparison. The water (blue), rock (red), and iron (green) curves are taken from \\citet{seag2007} and represent homogeneous solid planets. Intermediate compositions for differentiated planets are, from top down: dashed-blue, 75\\% H$_2$O, 22\\% MgSiO$_3$, and 3\\% Fe; dashed-dotted-blue, 48\\% H$_2$O, 48.5\\% MgSiO$_3$, and 6.5\\% Fe; dotted-blue, 25\\% H$_2$O, 52.5\\% MgSiO$_3$, and 22.5\\% Fe; dashed-red, 67.5\\% MgSiO$_3$ and 32.5\\% Fe; and dotted-red, 30\\% MgSiO$_3$ and 70\\% Fe. In general, the addition of a gas layer of up to $\\sim 5$\\% of the solid planet mass will inflate the radius of a rocky-iron planet through the range of radii corresponding to water planets with different water mass fractions. \\label{fig:gas1}} \\end{figure} As a second example, adding a gas layer of H/He equal to 1\\% of the mass of our fiducial planets increases the radius by $\\sim$ 20\\% of the original planet radius, or by about 0.35 $R_{\\oplus}$ for the planet masses we considered. Our major finding is that that exoplanets with a significant H/He layer cannot be distinguished from water worlds, based on $M_p$ and $R_p$ alone. For our fiducial solid exoplanets, adding up to 5\\% H/He by mass (for 10 $M_{\\oplus}$ planets) is sufficient to push the planet's radius through the entire range of radii corresponding to solid planets with no gas, including planets with up to 100 percent water composition. While we have not completed an exhaustive study of possible compositions, we find the non-uniqueness of water planets to be valid for any conditions we investigated. This generic finding holds for a wide range of assumptions of assumed temperatures. Taking our fiducial model, we vary $T_{eq}$, $T_{eff}$, and $\\gamma$ individually. For a 10 $M_{\\oplus}$ solid planet with an additional 0.1 $M_{\\oplus}$ H/He atmosphere, increasing $T_{eq}$ from 300 to 500 K increases the radius by about ~1\\% (Figure~2). For the same planet varying $T_{eff}$ from 10~K to 50~K results in an 8\\% increase in radius. While large, this value is comparable to expected radius uncertanties for these planets \\citep{gill2007a, gill2007b, demi2007}. Varying the altitude where radiation is absorbed (specified by $\\gamma$) has a much smaller effect on the planet radius. Varying $\\gamma$ from 0.1 to 10 causes the radius to decrease by 0.2\\%. \\begin{figure} \\plotone{f2} \\caption{The effects on the radius of varying the equilibrium temperature (top) and effective temperature (bottom) for a 10 $M_{\\oplus}$ planet with otherwise fiducial parameters. $T_{eq}$ values of 300, 400, and 500~K are plotted to simulate the effect of uncertainty in the orbital parameters and albedo on the expected radius. $T_{eff}$ values of 10, 30, and 50 K are plotted on the same scale as the $T_{eq}$ plot, to show uncertainties in the planet's interior temperature. Uncertainties in the internal energy of a planet lead to large variations in radii for a given mass, showing how temperature is a large uncertainty in the interpretation of a planet's internal composition. \\label{fig:temp}} \\end{figure} A corollary of our main result is that when a planet has a significant H/He atmosphere there is a wide degeneracy in allowable internal composition. This is not just compositional, but also relates to the trade off of temperature and mass of H/He gas. It could be argued that specifying a planet's composition implies a particular internal thermal profile derived from a consistent cooling history. As addressed in \\S2, the many unknowns and free input parameters for rocky planet interiors---such as the possible differences between atmospheric and interior compositions, equation of states, and the effect of tides on the planet's cooling history---prevent a self-consistent solution for the present time. How could a 5--20 $M_{\\oplus}$ exoplanet get a substantial H/He layer? Two different scenarios may produce them: direct capture of gas from the protoplanetary disk (possibly modified by the escape of some fraction of the original gas) or outgassing during accretion. A planet may capture and retain up to 1 to 2 $M_{\\oplus}$ of H/He if the planetary core did not grow quickly enough to capture more before the gas in the disk evaporated (as is the paradigm for Uranus and Neptune). Alternatively, for short-period exoplanets, a 1 to 2~$M_{\\oplus}$ H/He envelope may result after substantial loss of an initially massive gas envelope from irradiated evaporation \\citep[e.g.][]{bara2006}. \\citet{alib2006} consider atmospheric evaporation during migration, and conclude that the $10 M_{\\oplus}$ innermost planet in HD~69830, at 0.08 AU, kept $\\sim$2 $M_{\\oplus}$ of H/He over the 4 Gyr lifetime of the star. Little attention has been given to the mass and composition of exoplanet atmospheres from outgassing. Venus' atmosphere is $10^{-4} M_{\\oplus}$; if Venus had a surface gravity high enough to prevent H escape, its atmosphere would be over $10^{-3} M_{\\oplus}$. Even more massive H-rich atmospheres are possible. If a massive iron-silicate planet formed with enough water, the iron may react with the water during differentiation, liberating hydrogen gas \\citep{ring1979, waen1994}. L. Elkins-Tanton et al. (in prep.) estimate that the maximum H component is about six percent by mass for a terrestrial-composition planet. For a $10 M_{\\oplus}$ planet this would result in a $0.6 M_{\\oplus}$ H envelope. For short-period, low-mass planets, theoretical arguments of atmospheric escape may be the best way to identify a water world based on the mass and radius measurements alone \\citep{lege2004, sels2007}. Indeed, our assumption of H/He atmospheres for exoplanets relies on the condition that atmospheric mass loss has not evaporated all of the H/He. In the absence of hydrodynamic escape, the exospheric temperature (and not the atmospheric $T_{eff}$) drives the thermal Jeans escape of light gases. Earth and Jupiter both have exobase temperatures of 1000~K \\citep{depa2001}, significantly above their $T_{eff}$ of 255~K and 124.4~K respectively \\citep{cox2000}. Uranus and Neptune have exobase temperatures around 750~K \\citep{depa2001}. We note that because GJ 436 must have at least 1~$M_{\\oplus}$ of H/He, its exospheric temperature is not too high. On the other extreme, planets of $5 M_{\\oplus}$ would require very low exospheric temperatures ($\\sim 300$~K) to retain a massive atmosphere over the course of billions of years. Nevertheless, a young $5 M_{\\oplus}$ Earth-mass planet with a captured atmosphere could still have a H/He atmosphere and an old $5 M_{\\oplus}$ planet could retain a substantial He fraction, making its compositional identification ambiguous. \\begin{figure} \\plotone{f3} \\caption{Density vs. radius for three different potential compositions of GJ436b. From top to bottom: solid (black) curve, 19.0 $M_{\\oplus}$ core (30\\% Fe, 70\\% MgSiO$_3$) with 3.2 $M_{\\oplus}$ H/He (Y=0.28); dotted (red) curve, 20.0 $M_{\\oplus}$ core (100\\% (Mg,Fe)SiO$_3$) with 2.2 $M_{\\oplus}$ H (Y=0); dashed (blue) curve, 20.5 $M_{\\oplus}$ core (90\\% H$_2$O, 10\\% MgSiO$_3$) with 1.7 $M_{\\oplus}$ H/He (Y=0.28). All three planets have the same total radius (4.3 $R_{\\oplus}$) and total mass (22.2 $M_{\\oplus}$). \\label{fig:GJinterior}} \\end{figure} We now turn to a qualitative study of GJ436b, to show that the interpretation that GJ436b is a water world akin to Uranus and Neptune \\citep{gill2007b} is not the only possibility. We consider the GJ~436b values $M_p = 22.2 M_{\\oplus}$ and $R_p = 4.3 R_{\\oplus}$ from \\citet{demi2007}. The internal structure in Figure~3 shows how three planets with very different internal compositions can have the same total mass and radius. We first explore a planet similar to our fiducial model: a 22.2 $M_{\\oplus}$ solid planet with Earth-like iron/rock mass ratio (30/70), $T_{eff}=30$~K, and $T_{eq}=600$~K in rough agreement with the orbital parameters (assuming an albedo of 0.1). By adding $\\sim$3.2$M_{\\oplus}$ of hydrogen-helium to the 19.0 $M_{\\oplus}$ solid planet, we are able to reproduce GJ~436b's radius. We note that the mass of gas is 15\\% of the solid mass, likely too much to have originated from outgassing, and so capture must be at least partially invoked to explain such a massive atmosphere. The second composition for GJ~436b we considered is for water worlds, one with a 50\\% water mantle (by mass) and 50\\% silicate core, and another with 90\\% water mantle and a 10\\% silicate core. These planets also need some H/He to match the known radius, 12\\% and 8\\% by mass, respectively. The third model approximates planets with atmospheres created from outgassing, considering an extreme scenario where all of the available water has oxidized iron, leaving a 100\\% (Mg,Fe)SiO$_3$ solid planet core. To match the observed radius, a 22.2 $M_{\\oplus}$ planet requires $\\sim 2.2 M_{\\oplus}$ of H alone, a case that assumes no initial trapping and subsequent outgassing of He. The pure-hydrogen atmosphere is 10\\% of the mass of the solid planet. This is above the theoretical maximum of outgassing based on observed abundances of metallic iron in chondritic meteorites from our solar system (see Elkins-Tanton et al. in prep). Although not an exhaustive study, the range of interior compositions illustrates the variety of possibilities, though all models require some H/He. While our study is preliminary, we make the robust point that H-rich thick atmospheres will confuse the interpretation of planets based on a measured mass and radius. This point is independent of the uncertainties retained by our model including $T_{eq}$, $T_{eff}$, the mass fraction of H/He, and the mixing ratio of H and He. We find that the identification of water worlds based on the mass-radius relationship alone is impossible unless a significant gas layer can be ruled out by other means. Spectroscopy is the most likely means and may become routine with transit transmission and emission spectroscopy, and eventually with spectroscopy by direct imaging." }, "0710/0710.4036_arXiv.txt": { "abstract": "We present near-infrared (H- and K-band) integral-field observations of the circumnuclear star formation rings in five nearby spiral galaxies. The data, obtained at the {\\it Very Large Telescope} with the SINFONI spectrograph, are used to construct maps of various emission lines that reveal the individual star forming regions (\"hot spots\") delineating the rings. We derive the morphological parameters of the rings, and construct velocity fields of the stars and the emission line gas. We propose a qualitative, but robust, diagnostic for relative hot spot ages based on the intensity ratios of the emission lines \\brg , \\hel , and \\fetwo . Application of this diagnostic to the data presented here provides tentative support for a scenario in which star formation in the rings is triggered predominantly at two well-defined regions close to, and downstream from, the intersection of dust lanes along the bar with the inner Lindblad resonance. ", "introduction": "In many spiral galaxies of early- and intermediate Hubble type (Sa-Sc), active star formation is organized in a ring-like structure. These star formation rings offer a unique opportunity to study massive star formation in external galaxies: they produce up to 2/3 of the bolometric luminosity of their host galaxies \\citep[e.g. NGC\\,7469; ][]{gen95}, and often contain a large fraction of the entire star formation activity of the galaxy. The general picture of why molecular gas assembles in a ring is well understood as a natural consequence of a non-axisymmetric gravitational potential, nearly always due to the presence of a stellar bar or oval distortion \\citep[e.g.][]{com85,kna95,hel96,but96}. Because of its dissipative nature, molecular gas accumulates around the radii at which the stellar orbits experience dynamical resonances with the rotating bar potential. Depending on the pattern speed of the bar and the rotation curve of the galaxy, there can be one or multiple such resonances. The high gas densities found in the rings, combined with a variety of excitation mechanisms such as ultra-violet radiation from young stars and mechanical shocks due to energetic outflows from massive stars and/or an active galactic nucleus (AGN) help reveal the physical state of the interstellar matter (ISM), be it molecular, atomic, or ionized gas. Besides being fascinating laboratories in their own right, star formation rings are important also for the secular evolution of disk galaxies \\citep{kor04}. This is particularly true for the innermost of the dynamical resonances which is called either the \"inner Lindblad resonance\" (ILR) or - in cases where a compact massive object leads to an additional dynamical resonance - the ``nuclear Lindblad resonance'' \\citep*[NLR; ][]{fuk98}. Either of these resonances can produce gas rings with radii of a few hundred pc or less, depending on the enclosed mass, and the rotation speed of the galaxy disk. On these spatial scales, a number of processes can cause the gas to lose angular momentum and to subsequently flow towards the nucleus. Examples for such processes include torques due to the stellar potential \\citep{gar05}, dynamical friction between giant molecular clouds that form within the ring due to self-gravity of the gas \\citep*{fuk00}, the formation of spiral density waves \\citep{eng00}, or mechanical energy released by star formation in the ring via stellar winds and/or supernova explosions. Understanding the gas dynamics and star formation processes of nuclear\\footnote{Throughout this paper, the term ``nuclear ring'' is used to imply that it is the innermost (star forming) gas ring that can be resolved with the resolution of our data, typically a few hundred pc in diameter.} rings in disk galaxies is therefore crucial for developing models of gas accumulation at the very nucleus of a galaxy and the evolution of any compact massive object (CMO) which can exist in the form of a nuclear star cluster \\citep[NC,][]{boe02} and/or a supermassive black hole (SMBH). While some theoretical models exist on the gas behavior around a CMO \\citep[e.g.][]{fuk00}, observational data to constrain these models are rare, mostly because of the limited spatial resolution of mm-observations. Only recently has it become possible to study the molecular gas flows within a few tens of pc from the nucleus in a small number of nearby galaxies \\citep[e.g.][]{sch03,sch06,sch07}. In order to increase the number of well-studied nuclear rings, we have begun a project to study the near-infrared (NIR) properties of five such objects, using the SINFONI integral-field spectrograph on the Very Large Telescope (VLT). This paper discusses the morphologies, star formation rates, and kinematic properties of the rings. In a subsequent paper, we will make use of the spectroscopic information contained in the SINFONI data to investigate in more detail the star formation process, stellar populations, and gas excitation mechanism(s) in individual galaxies, both in the rings and the galaxy nuclei. This paper is structured as follows. In \\S\\ref{sec:data}, we describe the sample selection, observational details, and data reduction techniques common to all galaxies. The continuum and emission line morphology of the rings as well as the velocity fields of stars and gas for the individual galaxies are presented in \\S\\ref{sec:diagnostics}. We discuss the results in the context of competing models for the propagation of star formation in the rings in \\S\\ref{sec:discuss}, and summarize our analysis in \\S\\ref{sec:summary}. ", "conclusions": "} Based on high-resolution NIR integral-field observations of five nuclear star formation rings, we have presented their emission line morphologies and velocity structure both in gas and stars. We have introduced a new method to derive relative hot spot ages along the rings using the relative strengths of the \\hel , \\brg , and \\fetwo\\ lines. We employ this method to investigate the plausibilty of two competing scenarios for the way star formation is induced in nuclear rings, namely i) the ``popcorn'' model in which hot spots appear stochastically around the ring, and ii) the ``pearls on a string'' scenario in which star formation is triggered predominantly at two overdensity regions on either side of the ring. Only the latter predicts a well-ordered age sequence of hot spots along either half of a ring. The data presented in this study provide tentative support for the ``pearls on a string'' scenario, in that three out of five sample galaxies show some evidence for an age gradient of hot spots along the ring, while the remaining two galaxies have incomplete information and thus are consistent with either model. Of course, it might well be that star formation proceeds differently in some rings than in others. However, given that nuclear rings appear to form via a common mechanism (i.e. the gas response to a bar-shaped potential), it is not unreasonable to expect that they also follow a common path for inducing star formation. The small number of objects described here is clearly insufficient to provide reliable statistics, and similar studies of larger galaxy samples are needed to decide whether a particular mechanism governs the star formation in the majority of nuclear rings. Be that as it may, the proposed method of using NIR line ratios to estimate relative ages of young star clusters has been demonstrated to be a powerful tool for studies along this line." }, "0710/0710.5801_arXiv.txt": { "abstract": "We discuss the use of Sloan Digital Sky Survey (SDSS) {\\it ugriz} point-spread function (PSF) photometry for setting the zero points of {\\it UBVRI} CCD images. From a comparison with the Landolt (1992) standards and our own photometry we find that there is a fairly abrupt change in {\\it B, V, R, \\& I} zero points around $g, r, i \\sim 14.5$, and in the $U$ zero point at $u \\sim 16$. These changes correspond to where there is significant interpolation due to saturation in the SDSS PSF fluxes. There also seems to be another, much smaller systematic effect for stars with $g, r \\gtrsim 19.5$. The latter effect is consistent with a small Malmquist bias. Because of the difficulties with PSF fluxes of brighter stars, we recommend that comparisons of {\\it ugriz} and {\\it UBVRI} photometry should only be made for unsaturated stars with $g, r$ and $i$ in the range 14.5 -- 19.5, and $u$ in the range 16 -- 19.5. We give a prescription for setting the $UBVRI$ zero points for CCD images, and general equations for transforming from {\\it ugriz} to {\\it UBVRI}. ", "introduction": "When CCD images of a field are taken it is necessary to determine the photometric zero points from stars of known magnitudes. It is, however, not unusual for there to be no stars with $UBVRI$ photometry available. Fortunately, the Sloan Digital Sky Survey (SDSS) now provides homogenous $ugriz$ photometry for stars in a large fraction of the northern sky out of the plane of the Milky Way. Technical details of the SDSS are given in \\citep{york00} and \\citep{stoughton02}. The $ugriz$ system \\citep{fukugita96} is significantly different from the widely used $UBVRI$ Johnson-Cousins system \\citep{cousins76}, so it is necessary to transform between the two systems. A number of papers \\citep{fukugita96,smith02,karaali03,karaali05,bilir05,jordi06,rodgers06,ivezic07,davenport07,bilir07} have considered the transformations between $ugriz$ and $UBVRI$ (see Section 6 for a discussion of these transformations). During the course of using SDSS $ugriz$ photometry to establish the zero points for comparison stars for photometry of active galactic nuclei (AGN), we noticed that the zero points were different for the fainter stars in a field than for the brighter stars. The difference was in the sense that stars with $g \\lesssim 14$ were systematically brighter than predicted from the SDSS magnitudes. The difference did not seem to depend on the color of the stars and a check of the CCD used showed no evidence for non-linearity. A subsequent comparison of magnitudes of Landolt standards \\citep{landolt92} revealed a similar zero-point difference for stars brighter or fainter than $r \\sim 14$. In this note we report results of our investigation of the limitations of using SDSS photometry for bright stars, and give a prescription for setting zero points in CCD images taken through {\\it UBVRI} filters. ", "conclusions": "The transformations we give in equations (1) -- (5) are consistent with the range of previously published transformations. Because our linear transformation equations are derived for a practical purpose of calibrating $UBVRI$ photometry, and are available for each of the Johnson-Cousins filters individually, our transformations are different in nature from those previously published. We briefly summarize here previously published transformations and discuss how they differ from the ones given above. \\cite{fukugita96} give synthetic transformations from $UBVRI$ to $u'g'r'i'z'$. \\cite{smith02} gave transformations between $UBVRI$ and $u'g'r'i'z'$ magnitudes observed with the Photometric Telescope (PT) at Apache Point Observatory for some filters and for colors. \\cite{rodgers06} give improved color transformations between $u'g'r'i'z'$ and $UBVRI$ for main-sequence stars. They also consider higher-order color terms. It is important to note the difference between $u'g'r'i'z'$ and $ugriz$. This is discussed in \\cite{smith07}. Additional technical details concerning the difference between the two systems as well as transformations between them are discussed in \\cite{tucker06}. \\cite{jordi06} give color transformations between $ugriz$ as observed with the SDSS 2.5-m telescope (rather than the PT) and $UBVRI$. Additional transformations are given by \\cite{jester05}, \\cite{karaali03}, \\cite{karaali05},\\cite{bilir05}, \\cite{davenport07}, and \\cite{bilir07}. Some of the transformations including \\cite{ivezic07} consider polynomials in the color terms, but we found no need for higher-order terms for the restricted range of colors we consider. Note that the above cited transformations consider only colors, transform from $UBVRI$ to $ugriz$, are derived for the $u'g'r'i'z'$ system, or give transformations only for select Johnson-Cousins filters. In this note, our aim has been to give a practical means of photometrically calibrating $UBVRI$ CCD images. Researchers who are interested in astrophysical applications of SDSS photometry (such as the determination of spectroscopic parallaxes or fitting theoretical isochrones to HR diagrams) are referred to the above mentioned papers because the $ugriz$ to $UBVRI$ transformations depend on the luminosity class and metallicity of the stars. We have minimized these effects for zero-point setting by using a fairly tight color selection." }, "0710/0710.5030_arXiv.txt": { "abstract": "Optical high-resolution spectra of the R Coronae Borealis star V CrA at light maximum and during minimum light are discussed. Abundance analysis confirms previous results showing that V CrA has the composition of the small subclass of R Coronae Borealis (RCB) stars know as `minority' RCBs, i.e., the Si/Fe and S/Fe ratios are 100 times their solar values. A notable novel result for RCBs is the detection of the 1-0 Swan system $^{12}$C$^{13}$C bandhead indicating that $^{13}$C is abundant: spectrum synthesis shows that $^{12}$C/$^{13}$C is about 3 to 4. Absorption line profiles are variable at maximum light with some lines showing evidence of splitting by about 10 km s$^{-1}$. A spectrum obtained as the star was recovering from a deep minimum shows the presence of cool C$_2$ molecules with a rotational temperature of about 1200K, a temperature suggestive of gas in which carbon is condensing into soot. The presence of rapidly outflowing gas is shown by blue-shifted absorption components of the Na\\,{\\sc i} D and K\\,{\\sc i} 7698 \\AA\\ resonance lines. ", "introduction": "R Coronae Borealis stars (here, RCBs) are a rare class of peculiar variable stars. The two defining characteristics of RCBs are (i) a propensity to fade at unpredictable times by up to about 8 magnitudes as a result of obscuration by clouds of soot, and (ii) a supergiant-like atmosphere that is very H-deficient and He-rich. The subject of this paper, V Coronae Australis (V CrA) is even a peculiar member of this class of peculiar stars. It is a `minority' R CrB. The distinction between majority and minority members was made first by Lambert \\& Rao (1994) on the basis of an abundance analysis of warm RCBs. The minority RCBs are quite severely deficient in iron relative to the majority RCBs and to the Sun but some elements, particularly Si and S, have near-solar abundances in the minority (and majority) RCBs. This combination results in some very unusual abundance ratios, for example, the Si/Fe and S/Fe ratios of minority RCBs are approximately 100 times the solar ratios. V CrA also seems to be an especially lively producer of dust (Feast et al. 1997). The realization that V CrA is an unusual RCB led us to occasional spectroscopic monitoring of its optical spectrum. The current paper discusses high-resolution spectroscopic observations obtained on seven occasions between 1989 and 2003 when the star was either at or near maximum light or in decline by 3 to 5 magnitudes. A suitable spectrum at maximum light is subjected here to an abundance analysis. The previous analysis (Asplund et al. 2000) was based on a spectrum obtained during a shallow light minimum. This spectrum may have been contaminated by phenomena expected at minimum light (i.e., some lines may have been filled in partially by emission). In addition, our new spectra cover a broader bandpass at higher resolution and signal-to-noise ratio than the earlier spectrum. Other spectra show for the first time for V CrA line splitting indicative of the presence of an atmospheric shock. Spectra taken at minimum light are discussed in the context of our detailed studies of 1995-96 and 2003 minima of R CrB (Rao et al 1999; Rao, Lambert \\& Shetrone 2006). \\begin{figure} \\epsfxsize=8truecm \\epsffile{lcvcra1.ps} \\caption{The visual (red dots) light curve of V CrA showing the several light minima during 1988--1998 period. Dates on which four spectroscopic observations were obtained are indicated by a dashed line (also see Table 1). Upper limits to the brightness are shown by green unfilled circles. The observations are from the AAVSO database.} \\end{figure} ", "conclusions": "Differences between the spectra of RCBs at minimum light encourage us to continue our monitoring of V CrA and other RCBs. Studies of the initial stages of a decline should reveal clues to the trigger that sets off a decline. In this era when photometric observations by amateur astronomers are reported on the internet almost instantaneously, spectroscopic and other follow-up observations are limited by access to suitable telescopes. The advent of queue scheduling is smoothing the path to obtaining the follow-up observations. Development of a consortium of observers would also ease the situation. Observations of the RCBs in the deepest of minima are likely to provide novel data on their extended atmospheres, especially on the enigmatic broad lines which, the Na D lines apart, have been studied in detail with high-quality spectra only in the case of R CrB. For RCBs, aside from the three brightest stars R CrB, RY Sgr, and V854 Cen, a large telescope will be needed to acquire quality spectra at the faint magnitudes expected to reveal the broad lines." }, "0710/0710.2426_arXiv.txt": { "abstract": "Sgr A* is a source of strongly variable emission in several energy bands. It is generally agreed that this emission comes from the material surrounding the black hole which is either falling in or flowing out. The activity must be driven by accretion but the character of accretion flow in this object is an open question. We suggest that the inflow is dominated by the relatively low angular momentum material originating in one of the nearby group of stars. Such material flows in directly towards the black hole up to the distance of order of ten Schwarzschild radii or less, where it hits the angular momentum barrier which leads naturally to a flow variability. We study both the analytical and the numerical solutions for the flow dynamics, and we analyze the radiation spectra in both cases using the Monte Carlo code to simulate the synchrotron, bremsstrahlung and the Compton scattering. Our model roughly reproduces the broad band spectrum of Sgr A* and its variability if we allow for a small fraction of energy to be converted to non-thermal population of electrons. It is also consistent (for a range of viewing angles) with the strong constraints on the amount of circumnuclear material imposed by the measurements of the Faraday rotation. ", "introduction": "% The character of the accretion flow onto a black hole depends on the initial angular momentum of the material. This angular momentum is specified by the outer boundary conditions which depend on the relative motion of the donor with respect to the black hole. This angular momentum corresponds to a certain circularization radius, i.e. the radius where this angular momentum is equal to the local Keplerian value. In binary systems the material comes from the secondary star and in general is possess high angular momentum due to the orbital motion. In Low mass X-ray binaries the flow proceeds through an inner Lagrange point and the circularization radius is a significant fraction of a Roche radius around a black hole, of order of $10^4 R_g$ ($R_g = GM/c$). In high mass X-ray binaries the accretion flow comes from the intercepted focused wind, so the circularization radius is smaller but still large, of order of $10^3 R_g$. In such case the inflowing material form an accretion disk around a black hole and the inflow proceeds due to the angular momentum transfer. Apart from the outermost region and the region close to the ISCO (innermost stable circular orbit), the distribution of the angular momentum is relatively smooth and not much different from the Keplerian law. The exact departures from Keplerian motion depends on the disk temperature (or more exactly, on the pressure distribution). In active galactic nuclei (AGN) the source of material is less specified. The material comes either from the stars (in the form of stellar winds) or from the gaseous phase of the galactic material. Bright AGN (quasars, Seyfert 1 galaxies) show the presence of accretion disks similar to the disks in binary systems so we can conclude that the angular momentum reaching the galactic center is high. In sources showing water maser activity we observe the outer parts of the disk directly, and in most sources the motion of the disk material is Keplerian. However, in weakly active galaxies like Sgr A* or giant elliptical galaxies we see no direct evidence of a disk. In Sgr A* the presence of the cold disk is actually excluded by the lack of eclipses of the stars which move very close to the central black hole and are systematically monitored since several years. Since in weakly active galaxies there are no direct observational arguments for any value of the angular momentum of the donated material and the location of material sources, three types of models are being considered: \\begin{itemize} \\item high angular momentum flow, with circularization radius of order of hundreds-thousands of $R_g$ \\item low angular momentum flow, with circularization radius of order of a few $R_g$ \\item spherical and quasi-spherical accretion, without angular momentum barrier. \\end{itemize} The high angular momentum flow solutions for weakly active galaxies generally belong to ADAF (advection dominated accretion flow) family \\citep{1977ApJ...214..840I,1994ApJ...428L..13N}, with possibly additional effects like outflows \\citep{1999MNRAS.303L...1B} and convection. In this case the flow is not exactly Keplerian since the pressure gradients are important, but the local ratio of the angular momentum to the Keplerian angular momentum in most part of the flow is not wildly different from unity, and the angular momentum transfer (through viscosity) or angular momentum loss (through magnetic wind) at all radii is essential. Stationary solutions usually exist, and asymptotically the density of the flow approaches zero at infinity. In spherical and quasi-spherical flow there is no angular momentum barrier so the loss of angular momentum is not the necessary condition for the accretion to occur. Examples of such solutions are: purely spherical Bondi flow or flows where the angular momentum density is below the minimum angular momentum at the circular orbit around a black hole which is given by \\begin{equation} l_{min} = 3 \\sqrt{3} GM/c \\end{equation} in case of Schwarzschild black hole; more general formula for a Kerr black hole can be found in \\citep{1972ApJ...178..347B}. In Bondi solution \\citep{1952MNRAS.112..195B,2003ApJ...591..891B} the outer boundary condition are specified by the density and the temperature of the uniform medium surrounding black hole at large distances. The flow velocity is zero at infinity, the inflow becomes transonic at the Bondi radius, and the supersonic flow reaches the black hole horizon. The Bondi radius depends significantly on the gas properties (e.g. politropic index; Bondi radius is of order of thousands of $R_g$ for relativistic flow with $\\gamma = 4/3$ but is approaches zero if $\\gamma \\rightarrow 5/3$, typical for perfect fluid non-relativistic solution), but the accretion rate is much less sensitive to those assumptions.Purely Bondi flow has generally very low radiative efficiency so it cannot reproduce the observed luminosity in most weakly active galaxies \\citep{2006A&A...450...93M}. If the accreting material at the outer boundary condition has certain angular momentum $l < l_{min}$, the dynamics of the flow is slightly modified in comparison with Bondi flow and the flow is not spherically symmetric any more but the stationary solution for the flow always exists. The intermediate case of low angular momentum the situation is the most complex as initially the flow behaves as the Bondi flow but close to the black hole the flow starts suddenly to feel the angular momentum barrier \\citep{1981ApJ...246..314A}. In this case analytical stationary solutions frequently do not exist. In numerical solutions the flow is variable and does not reach a stationary solution in the computing time. If the angular momentum of the donated material is also a subject of changes (e.g. the result of the stellar motion), a truly stationary solution indeed can never be reached for physical reasons. In the case of Sgr A* the available spatial resolution is the highest and we can have the best insight into the sources of material \\citep{2007IAUS..238..173G}. Therefore, in the present paper we concentrate specifically on this source and we argue that the low angular momentum flow is an interesting and promising option for the flow description. ", "conclusions": "% Low angular momentum accretion flow is a promising scenario for the accretion onto Sgr A* due to its natural variability pattern. The flow is slightly more energetically efficient than the purely spherical Bondi flow and can reproduce both the required level of the luminosity and is consistent with the data on Faraday rotation measure. The overall broad band spectra are also roughly reproduced if a fraction of energy is allowed to be converted the non-thermal population of electrons. The current results are therefore encouraging, and the further work is in progress. \\ack% The present work was supported by the Polish Grant 1P03D~008~29 and the Polish Astroparticle Network 621/E-78/SN-0068/2007." }, "0710/0710.5354_arXiv.txt": { "abstract": "We simulate the collisional formation of a ring galaxy and we integrate its evolution up to 1.5 Gyr after the interaction. About $100-200$ Myr after the collision, the simulated galaxy is very similar to observed ring galaxies (e.g. Cartwheel). After this stage, the ring keeps expanding and fades. Approximately $0.5-1$ Gyr after the interaction, the disc becomes very large ($\\sim{}100$ kpc) and flat. Such extended discs have been observed only in giant low surface brightness galaxies (GLSBs). We compare various properties of our simulated galaxies (surface brightness profile, morphology, HI spectrum and rotation curve) with the observations of four well-known GLSBs (UGC6614, Malin 1, Malin 2 and NGC7589). The simulations match quite well the observations, suggesting that ring galaxies could be the progenitors of GLSBs. This result is crucial for the cold dark matter (CDM) model, as it was very difficult, so far, to explain the formation of GLSBs within the CDM scenario. ", "introduction": "Ring galaxies are one of the most intriguing categories of peculiar galaxies. About 280 galaxies have been classified as ring-like in the Catalogue of Southern Peculiar Galaxies and Associations (CPGA; Arp \\& Madore 1987). They have commonly been divided in two different classes, P-type and O-type ring galaxies (Few \\& Madore 1986). The former consists of galaxies where the nucleus is often off-centre and the ring is quite knotty and irregular, while the latter includes objects where the nucleus is central and the ring is regular. Even if for some objects such classification is ambiguous, the number of P-type and of O-type ring galaxies are roughly comparable. The differences among the two classes are probably connected with the formation mechanism of such galaxies: for P-type ring galaxies a collisional origin has been proposed (Lynds \\& Toomre 1976; Theys \\& Spiegel 1976; Appleton \\& Struck-Marcell 1987a, 1987b; Hernquist \\& Weil 1993; Mihos \\& Hernquist 1994; Appleton \\& Struck-Marcell 1996; Struck 1997; Horellou \\& Combes 2001), as most of them have at least one nearby companion (Few \\& Madore 1986); whereas O-type ring galaxies can be resonant (R)S galaxies (de Vacouleurs 1959). Simulations of galaxy collisions leading to the formation of P-type ring galaxies show that the ring phase is quite short-lived (Hernquist \\& Weil 1993; Mihos \\& Hernquist 1994; Horellou \\& Combes 2001; Mapelli et al. 2007, hereafter M07): the simulated disc galaxy develops a ring similar to the observed ones $\\approx{}100$ Myr after the collision with the intruder; but the ring remains dense and clearly visible only for the first $\\approx{}300$ Myr (M07). Thus, ring galaxies are expected to rapidly evolve into something else. Up to now, only analytic models for ring waves have been adopted to study the late stages of the ring galaxy life (Struck-Marcell \\& Lotan 1990; Appleton \\& Struck-Marcell 1996). Neither observations nor simulations have been carried on to investigate the fate of ring galaxies after the ring phase, leaving a lot of uncertainties. This paper aims at describing the subsequent stages of the evolution of a ring galaxy (up to $\\sim{}1.5$ Gyr), by means of SPH/$N$-body simulations. Our simulations suggest a possible link between old ring galaxies and the so-called giant low surface brightness galaxies (GLSBs). The GLSBs are low surface brightness galaxies (LSBs) characterized by unusually large extension of the stellar and gaseous discs (up to $\\sim{}100$ kpc; Bothun et al. 1987; Impey \\& Bothun 1989; Bothun et al. 1990; Sprayberry et al. 1995; Pickering et al. 1997; Moore \\& Parker 2007) and (often) by the presence of a normal stellar bulge (Sprayberry et al. 1995; Pickering et al. 1997). The existence of GLSBs has always been puzzling. Galaxy formation simulations within the cold dark matter (CDM) model have serious difficulties in producing realistic disc galaxies. In such simulations, too much angular momentum is lost during the assembly of objects, producing discs that tend to be too compact and dense. While a 'Milky Way-like' galaxy can be reproduced by high-resolution CDM simulations (provided that its merging history is fairly quiet and that heating by supernovae is properly accounted for; Governato et al. 2007), the extended discs of GLSBs require that much more angular momentum is preserved during the hierarchical build up. Such extended discs are thus beyond the reach of current galaxy formation simulations, and it is unclear whether improving the realism of such simulations will solve the problem. Hoffman, Silk \\& Wyse (1992) proposed that GLSBs form from rare density peaks in voids. However, most of currently known GLSBs do not appear to be connected with voids and often show interacting companions (Pickering et al. 1997). A more promising scenario is the formation of GLSBs from massive disc galaxies due to a bar instability: a large scale bar can redistribute the disc matter and significantly increase the disc scale-length (Noguchi 2001; Mayer \\& Wadsley 2004). In this case the redistribution of angular momentum by the bar instability could counteract the natural tendency of hierarchical assembly to remove angular momentum from the disc material (Kaufmann et al. 2007). Bar instabilities, however, normally do not increase the disc scale length by more than a factor of 2-2.5 (Debattista et al. 2006; Kaufmann et al. 2007). In this paper we show that also the propagation of the ring in an old collisional ring galaxy can lead to the redistribution of mass and angular momentum in both the stellar and gas component out to a distance of $\\sim{}100-150$ kpc from the centre of the galaxy, producing features (e.g. the surface brightness profile, the star formation, the HI emission spectra and the rotation curve) which are typical of GLSBs. ", "conclusions": "In this paper we presented a numerical model of ring galaxy evolution. About $100-200$ Myr after the collision with the intruder, the target disc galaxy evolves into a ring galaxy, similar to Cartwheel. Afterward, the ring progressively expands and fades. After $\\approx{}0.5-1.0$ Gyr the ring is no longer distinguishable from the disc, its surface density is more than 1 order of magnitude lower than in the Cartwheel phase, and the disc extends up to $\\sim{}100$ kpc. We showed that simulated ring galaxies in the late stages of their dynamical evolution ($\\gtrsim{}500$ Myr) are very similar to the observed GLSBs. The $R$-band surface brightness profile, the SFR, the HI spectra and also the rotation curves of four GLSBs are well reproduced by the simulations. This result is unlikely due to a simple coincidence. If all GLSBs were originated by the evolution of P-type ring galaxies, their current number density would be comparable to the observed number density of P-type ring galaxies, i.e. $\\sim{}5.4\\times{}10^{-6}\\,{}h^3\\,{}{\\rm Mpc}^{-3}$ (where $h$ is the Hubble constant; Few \\& Madore 1986). Since we know $\\sim{}17$ galaxies which can be classified as GLSBs (Sprayberry et al. 1995; Pickering et al. 1997) and which are within $\\sim{}340$ Mpc from our Galaxy, the lower limit of the current number density of GLSBs is $\\sim{}2.7\\times{}10^{-7}\\,{}h^3\\,{}{\\rm Mpc}^{-3}$, i.e. about one order of magnitude lower than the density of ring galaxies. This can imply either that a large fraction of GLSBs have not been detected yet, or that only the $\\sim{}$5 per cent of ring galaxies ends up into a GLSB. The former scenario is quite realistic, as magnitude-limited surveys are strongly biased against GLSBs (Bothun et al. 1997). The latter hypothesis is also likely, as ring galaxies can end their life in other ways. For example, if the intruder is not sufficiently massive, the density wave is not strong enough to produce a GLSB, or, if the relative velocity is not sufficiently high, the companion can come back and merge or disrupt the propagating wave. Furthermore, recycled dwarf galaxies might also form from the debris of old collisional ring galaxies (e.g. the case of NGC~5291; Bournaud et al. 2007). Furthermore, our model can explain the formation of GLSBs within the context of the CDM scenario, in which very extended discs are hardly explained as a result of normal hierarchical assembly. Hence it will be quite important to check the consistency of our model with future observations. For example, the available HI data (Pickering et al. 1997) show that most of GLSBs have strong non-circular motions. From the published data it is not possible to understand whether these motions are consistent with the expansion and/or the falling back of gas in the ring. Thus, it is very important to make new radio observations of GLSBs or, at least, re-examine the archival data. Another important point to address is how many GLSBs have a nearby companion which could be the intruder. Among the 4 GLSBs considered here, NGC~7589 has a well-known interacting companion (Pickering et al. 1997). Furthermore, UGC~6614 is surrounded by other galaxies with approximately the same redshift (e.g. KUG 1136+173, AGC~211143 and CGCG~097$-$034), but no studies have been done to establish whether they are in the same group as UGC6614. Finally, the surroundings of Malin~1 are populated by many companion candidates, but no redshift measurements are available at present (Sprayberry et al. 1995). A further issue raised by this paper is whether there are objects in the intermediate stage between ring galaxies and GLSBs. The ring galaxy Arp~10 has been thought to be a relatively old ring galaxy, because its rings are not as prominent as those of other ring galaxies (e.g. Cartwheel) and because its SFR, from H$\\alpha{}$ observations, appears quite low (Charmandaris, Appleton \\& Marston 1993). However, Bizyaev, Moiseev \\& Vorobyov (2007) have shown that the SFR of Arp~10, from far-infrared observations , is $\\sim{}10-21\\,{}M_\\odot{}$ yr$^{-1}$ (analogous to the one of Cartwheel; Mayya et al. 2005), and that the collision which produced the ring occurred only $\\sim{}85$ Myr ago. Then, both current observational data and theoretical models are not sufficient to support the idea that Arp~10 is an old ring galaxy\\footnote{Even our simulations cannot solve the problem, as they indicate that the surface brightness profile of Arp~10 can be matched by a $\\sim{}80$ Myr old ring galaxy (in agreement with Bizyaev et al. 2007), as well as by a $\\sim{}300$ Myr old ring galaxy. In particular, the surface brightness profile of Arp~10 is matched by a $80$ Myr old ring galaxy with the same initial conditions as run A but with $R_d=8.8$ kpc. Nice agreement with the observations is obtained also for a $300$ Myr old ring galaxy with the same initial conditions as runs B in table 1 of M07, but with $R_d=13.2$ kpc.}. On the other hand, one of the four GLSBs considered in this paper, UGC6614, shows (both in H$\\alpha{}$, in HI and in optical) structures which look like the remnants of the inner ring. Thus, UGC6614 could be in an intermediate, connecting stage between GLSBs and ring galaxies. It would be crucial to study more deeply UGC6614, as well as to search for other galaxies which could represent the intermediate phase between GLSBs and ring galaxies." }, "0710/0710.5162_arXiv.txt": { "abstract": "We present recent results from a Keck study of the composition of the Galactic bulge, as well as results from the bulge Bulge Radial Velocity Assay (BRAVA). Culminating a 10 year investigation, Fulbright, McWilliam, \\& Rich (2006, 2007) solved the problem of deriving the iron abundance in the Galactic bulge, and find enhanced alpha element abundances, consistent with the earlier work of McWilliam \\& Rich (1994). We also report on a radial velocity survey of {\\sl 2MASS}-selected M giant stars in the Galactic bulge, observed with the CTIO 4m Hydra multi-object spectrograph. This program is to test dynamical models of the bulge and to search for and map any dynamically cold substructure in the Galactic bulge. We show initial results on fields at $-10^{\\circ} < l <+10^{\\circ}$ and $b=-4^{\\circ}$. We construct a longitude-velocity plot for the bulge stars and the model data, and find that contrary to previous studies, the bulge does not rotate as a solid body; from $-5^{\\circ}$14.6 \\\\ 19$^{\\rm h}$ 29$^{\\rm m}$ 0\\fs70 & 9\\degr{} 38\\arcmin{} 44\\farcs78 & 20.6 & 19.5 & $>$14.4 \\\\ \\hline \\end{tabular} \\end{table} \\pars{} is situated close to the Galactic plane ($l=45.8^{\\circ}$, $b=-3.8^{\\circ}$), in a molecular cloud called Cloud\\,A. This cloud was identified by \\citet{dame85} in their CO survey of molecular clouds in the northern Milky Way. The cloud occupies an area of 8 square degrees in the sky, and the only known young star associated with it is the T\\,Tauri star AS\\,353 \\citep{dame85}. In order to check whether FUors are usually associated with star forming regions, we searched the literature and found that most FUors are located in areas of active star formation \\citep[e.g.][]{henning}. To find out whether there is star formation in the vicinity of \\pars{}, we searched for pre-main sequence stars. For this purpose we constructed a 2MASS J$-$H vs.~H$-$K$_S$ and an IRAC [3.6]$-$[4.5] vs.~[5.8]$-$[8.0] colour-colour diagram for sources found in our $5\\farcm3\\,{\\times}\\,5\\farcm3$ IRAC field of view (Fig.~\\ref{fig:cc2mass}, \\ref{fig:ccirac}). Our selection criteria in the case of 2MASS was S/N ${>}\\,10$ and uncertainties ${<}\\,0.1\\,$mag in all J, H and K$_S$ bands, while in the case of IRAC S/N ${>}\\,3$ and detectability at all four bands were required. The 2MASS diagram revealed that most of the nearby objects are reddened main sequence or giant stars. On the IRAC diagram, however, there are three objects (apart from \\pars{} itself), which display infrared excess at $8\\,\\mu$m (marked by A, B and C in Fig.~\\ref{fig:ccirac}). According to the classification of \\citet{allen}, Class II sources exhibit colours of [3.6]$-$[4.5] ${>}\\,0.0$ and [5.8]$-$[8.0] ${>}\\,0.4$, thus one of these stars (A) might be a Class II source, while B and C are more likely Class III/main sequence sources. The nature of source A and its possible relationship to \\pars{} is yet to be investigated. Nevertheless, \\pars{} seems to be rather isolated compared to most FUors and certainly not associated with any rich cluster of young stellar objects. We also searched for possible close companions of \\pars{} in the WFPC2 and NACO direct images. In order to establish a detection limit for source detection, we measured the sky brightness on the NACO images (before sky-subtraction), and estimated a limiting magnitude for each filter. The resulting values are $22.8$, $21.6$, and $15.2\\,$mag in H, K$_S$ and L$^{\\prime}$, respectively. In case of the HST/WFPC2 image, the larger ($80\\,{\\times}\\,80$ arcsec) field of view made it possible to estimate a limiting magnitude using star counts; the resulting value is $23.5\\,$mag. Due to the bright reflection nebula, the detection limit is somewhat lower close to the star. The two closest objects we found are the following: one star to the southeast, at a distance of $1.4$ arcsec (560\\,AU at 400\\,pc), and another one to the northwest, at a distance of $3.3$ arcsec (1320\\,AU at 400\\,pc). These sources are marked with arrows in Fig.~\\ref{fig:images}. Neither of the stars are visible at $3.8$ or $0.814\\,\\mu$m, although we can give an upper limit for their L$^{\\prime}$ brightness. Their positions and photometry are given in Table~\\ref{tab:twostars}. As these sources are very red, they can equally be heavily reddened background stars, or stars with infrared excess (indicating that they might be associated with \\pars{}). Supposing that they are reddened main sequence stars, one can estimate an extinction of A$_V\\,{\\approx}\\,10\\,{-}\\,15\\,$mag. Further multifilter observations may help to clarify the nature of these objects and their possible relationship to \\pars{}. \\subsection{The circumstellar environment of \\pars{}} \\begin{figure} \\centering \\includegraphics[angle=0,width=0.45\\linewidth]{kospal_fig12.ps} \\caption{Sketch of the morphology of circumstellar material around \\pars{}, overlaid on the HST/WFPC2 image. The central star is surrounded by an edge-on disc. Perpendicular to the disc, the star drives a bipolar outflow that excavates an outflow cavity in the dense circumstellar material. Light from the central star illuminates the walls of the cavity.} \\label{fig:morph} \\end{figure} \\begin{figure} \\centering \\includegraphics[angle=90,width=0.95\\linewidth]{kospal_fig13.ps} \\caption{Brightness profiles of \\pars{} at $0.8\\,\\mu$m and in the H and K$_S$ bands. Dotted lines mark Hubble's relation.} \\label{fig:metszet} \\end{figure} \\begin{figure} \\centering \\includegraphics[angle=0,width=0.95\\linewidth]{kospal_fig14.ps} \\caption{South-north cut at $0.6$ arcsec east from the star. {\\it Solid line:} total intensity; {\\it dashed line:} polarized intensity.} \\label{fig:cutok} \\end{figure} \\begin{figure*} \\centering \\includegraphics[angle=90,width=0.95\\linewidth]{kospal_fig15.ps} \\caption{VLT/NACO polarization map of \\pars{} overlaid on H-band total intensity contours. The circle in the upper right corner displays the FWHM of the polarization measurement. Polarization vectors are displayed at full resolution, only showing the central low polarization band, where the polarization vectors are aligned. Such arrangement is expected when multiple scattering occurs in an edge-on disc.} \\label{fig:kiskep} \\end{figure*} The appearance and polarization properties of the nebula around \\pars{} can be understood in the following way: the star drives an approximately north-south oriented bipolar outflow, which had excavated a conical cavity in the dense circumstellar material (Fig.~\\ref{fig:morph}). The star illuminates this cavity and the light is scattered towards us mainly from the walls of the cavity. The outflow direction is perpendicular to an almost edge-on dense circumstellar disc. This picture is supported by the following facts: (a) the centrosymmetric polarization pattern is characteristic of reflection nebulae with single scattering; (b) the morphology and limb brightening suggest a hollow cavity (as opposed to an ``outflow nebula'', where the lobes are composed of dense material ejected by the central source); and (c) the low-polarization lane across the star strongly suggests the presence of an edge-on circumstellar disc, where multiple scattering occurs. In general the \\pars{} system shows similarities to the NGC\\,2261 nebula associated with R\\,Mon. This object also consists of a northern cometary nebula and a southern jetlike feature \\citep{warren}. \\subsubsection{Envelope/Cavity} \\label{sec:env} We characterised the opening of the upper lobe by marking the ridge along the northeastern and northwestern arcs which we interpret as the walls of the cavity. As viewed from the star northwards, the cavity starts as a cone with an opening angle of ${\\approx}\\,60^{\\circ}$, giving the nebula in Fig.~\\ref{fig:images} a characteristic equilateral triangle-shape. Farther away from the star the cavity deviates from the conical shape, becomes narrower. The whole cavity occupies an area of $8\\,000\\,{\\times}\\,24\\,000\\,$AU (at a distance of $400\\,$pc). The sharp outer boundary of the nebula implies a significant density contrast between the cavity and the surrounding envelope. In Fig.~\\ref{fig:metszet} we plotted radial brightness profiles at $0.8\\,\\mu$m and in the H and K$_S$ bands, starting from the star northwards. The slope of the intensity profiles in the inner ${\\sim}\\,1.6''$ ($640\\,$AU) follows closely Hubble's relation \\citep{hubble}, i.e.~the brightness is proportional to r$^{-2}$. This trend can be clearly followed above the noise out to about $5\\,$arcsec in our H and K$_S$ images, giving an estimate of $2000\\,$AU for the outer size of the envelope. The fact that the brightness profiles follow Hubble's relation implies that the nebula is produced by isotropic single scattering, in accordance with the centrosymmetric polarization pattern and the high degree of polarization. Around ${\\sim}\\,1.6''$ the profiles becomes steeper, probably due to decreased density in the inside of the cavity. Similar steepening in the outer part of the nebula was already mentioned by \\citet{li}. The profiles are similar at all observed wavelengths and do not show significant dependence on the position angle. \\subsubsection{Disc} \\label{sec:disc} As mentioned in Sect.~\\ref{sec:pol}, both the NACO and NICMOS polarimetric images show a lane of low polarization oriented nearly east-west across the star (Fig.~\\ref{fig:pol}). The drop in the degree of polarization can also be clearly seen in Fig.~\\ref{fig:cutok}, where a north-south cut at $0\\farcs6$ east to the star is plotted. Following \\citet{bm90}, we interpret the low polarization by multiple scattering in an edge-on disc (possible other explanations for the origin of the low polarization areas are discussed in e.g.~\\citealt{lr}). According to models of such circumstellar structures \\citep[e.g.][]{whitney, fischer} the polarization vectors are oriented parallel to the disc plane. As can be seen in Fig.~\\ref{fig:kiskep}, despite the low degree of polarization, the predicted alignment of the vectors can be clearly seen in the case of \\pars{} too. The NICMOS polarization map shows a similar effect. The dark lane in Fig.~\\ref{fig:pol} can be followed inwards to as close as $48\\,$AU from the star. This is an upper limit for the inner radius of the circumstellar disc. An interesting feature of the mentioned models is two depolarized areas on either side of the central star, which mark the outer end points of an edge-on disc. The depolarization is due to a transition from the linearly aligned to the centro-symmetic polarization pattern. These depolarized areas can be seen in Fig.~\\ref{fig:vector} bottom, on both sides at about $0.9''$ from the star. At the distance of \\pars{} this corresponds to $360\\,$AU, and can be adopted as the outer radius of the dense part of the disc, where multiple scattering at near-infrared wavelengths is dominant. It is interesting that the NACO map does not show clear depolarized areas at the same position, but seems to resolve the transition in vector orientation, while the larger beam of NICMOS averaged the differently oriented vectors, resulting in depolarized spots. The polarized intensity and degree of polarization images in Fig.~\\ref{fig:pol} suggest a slight asymmetry in the dense disc: the eastern (left) side is straight, while the western (right) side shows a kink and also has a different position angle than that on the other side. The thickness of the disc can be measured on the area where the polarization vectors are aligned (Fig.~\\ref{fig:kiskep}). This approximately corresponds to the area where the degree of polarization is below ${\\approx}\\,10\\%$. The resulting thickness is approximately $0.1''$ ($40\\,$AU) to the east and is somewhat larger, $0.2''$ ($80\\,$AU), to the west. One should note that, since these values are close to the spatial resolution of the polarimetric images, these numbers should be considered as upper limits for the thickness of the circumstellar disc around \\pars{}. They are upper limits also because if the inclination is not exactly 90$^{\\circ}$, the thickness of the disc can be even less. The thickness does not show significant increase with radial distance, suggesting the picture of a flat, rather than a flared disc, at least considering the dense, multiple-scattering part. The smooth brightness and polarization distribution between the disc and the surrounding envelope (Fig.~\\ref{fig:cutok}), however, implies that there is a continuous density transition between the two components. From the ratio of the horizontal to vertical sizes of the disc a lower limit of $84^{\\circ}$ for the inclination of the system (the angle between the normal of the disc and the line of sight) can be derived. \\citet{fischer} computed a grid of polarization maps of young stellar objects with the aim of helping the interpretation of polarimetric imaging observations. They consider five different models, four with massive, self-gravitating discs and one with a massless Keplerian disc. Since the mass of circumstellar material of \\pars{} derived from submillimetre observations is relatively low (${<}\\,0.3\\,$M$_{\\odot}$, \\citealt{henning, polom, sw, hillen}), the most appropriate model for our case is a Keplerian disc. Indeed the polarization pattern as computed by \\citet{fischer} for an inclination of 87$^{\\circ}$ (their Fig.~1) looks remarkably similar to our Fig.~\\ref{fig:vector} (the differences might be explained by the narrower cavity and flatter disc of \\pars{}). Thus, the geometry and structure assumed by \\citet{fischer} in their Keplerian model could be a good starting point for further radiative transfer modelling of \\pars{}. \\subsection{Modelling the circumstellar environment} \\label{sec:modelling} \\begin{table} \\centering \\caption{Model parameters. Parameters in italics are fixed, while the others were fitted.} \\label{tab:par} \\begin{tabular}{lcc} \\hline Parameter & Variable & Value \\\\ \\hline Inner disc radius & $R_1$ & 3.5\\,$R_\\odot$ \\\\ {\\it Outer disc radius} & $R_2$ & {\\it 360\\,AU} \\\\ Temperature at 1 AU & $T_{\\rm d,0}$ & 285\\,K \\\\ {\\it Power-law index for temperature} & $q_{\\rm d}$ & {\\it 0.75} \\\\ Power-law index for surface density & $p_{\\rm d}$ & 1.6 \\\\ {\\it Disc mass} & $M_{\\rm d}$ & {\\it 0.02\\,M$_\\odot$} \\\\ {\\it Inclination} & $i$ & {\\it 86$^\\circ$} \\\\ Inner envelope radius & $R_3$ & 5.4\\,AU \\\\ {\\it Outer envelope radius} & $R_4$ & {\\it 2000\\,AU} \\\\ Temperature at 5 AU & $T_{\\rm e,0}$ & 368\\,K \\\\ {\\it Power-law index for temperature} & $q_{\\rm e}$ & {\\it 0.4} \\\\ Power-law index for surface density & $p_{\\rm e}$ & 0.4 \\\\ Envelope mass & $M_{\\rm e}$ & 0.02\\,M$_\\odot$ \\\\ {\\it Interstellar Extinction} & $A_{V}$ & {\\it 2\\,mag}\\\\ \\hline \\end{tabular} \\end{table} Our observations provide some direct measurements of the geometry of the circumstellar structure (disc size and thickness, inclination, envelope size). In the following we discuss the consistency of this picture with the observed SED. Our approach is to construct a simple disc+envelope model, in which we fix those parameters whose values are known from our NACO observations or from other sources (outer disc radius from this work; power-law index for disc temperature from \\citealt{shakura}; disc mass was set in order to ensure that the whole disc is optically thick; inclination from this work; outer envelope radius from this work, the power-law index for envelope temperature is a typical value for optically thin envelopes containing larger than interstellar grains, e.g.~\\citealt{hartmannkonyv}, Eqn.~4.13; interstellar extinction from \\citealt{hillen}). Then we check whether the SED can be fitted by tuning the remaining parameters. We adopted an analytical disk model \\citep{adams}, which has been successfully used to model FUors \\citep{quanz2, v1647ori_midi}. Our model consists of two components, an optically thick and geometrically thin accretion disc \\citep{shakura} and an optically thin envelope (no cavity is assumed). No central star is included in the simulation, partly because in outbursting FUors the star's contribution is negligible compared to that of the inner disc \\citep{hk96}, and partly because of the edge-on geometry where the star is obscured by the disc. This assumption is supported by the fact that the shape of the SED at optical wavelengths is broader than a stellar photosphere. The model also does not take into account internal extinction and light scattering, thus it cannot reproduce any of the near-IR imaging and polarimetric observations. The temperature and surface density distribution in the disc are described by power-laws: \\begin{equation} T(r)=T_{{\\rm d}, 0}\\left(\\frac{r}{1\\,{\\rm AU}}\\right)^{-q_{\\rm d}}, \\end{equation} \\begin{equation} \\Sigma(r)=\\Sigma_{{\\rm d},0}\\left(\\frac{r}{1\\,{\\rm AU}}\\right)^{-p_{\\rm d}}. \\end{equation} Similar power-laws were assumed for the envelope. The observed flux at a specific frequency is given by \\begin{eqnarray} F_\\nu &=& \\frac{\\cos{i}}{D^2}\\int_{\\rm R1}^{\\rm R2}2\\pi r(1-e^{\\frac{-\\Sigma_{{\\rm d}}\\kappa_\\nu}{\\cos{i}}})B_\\nu(T_{\\rm d}){\\rm d}r + \\\\ & & \\frac{1}{D^2}\\int_{\\rm R3}^{\\rm R4}2\\pi r(1-e^{-\\Sigma_{{\\rm e}}\\kappa_\\nu})B_\\nu(T_{{\\rm e}}){\\rm d}r. \\end{eqnarray} The first term describes the emission of the accretion disc, the second term describes the radiation of the optically thin envelope. For the dust opacity we used a constant value of $\\kappa_{\\nu}\\,{=}$ 1 cm$^2$g$^{-1}$ at $\\lambda\\,{>}\\,1300\\,\\mu$m, $\\kappa_{\\nu}\\,{=}\\,\\kappa_{1300\\mu\\rm m}\\left(\\frac{\\lambda}{1300\\mu\\rm m}\\right)^{-1}$ between $1300$ and $100\\,\\mu$m and again a constant value of $\\kappa_{\\nu}\\,{=}\\,\\kappa(100\\,\\mu\\rm m)$ at $\\lambda\\,{<}\\,100\\,\\mu$m. We fitted the SED via $\\chi^2$ minimisation using a genetic optimization algorithm PIKAIA \\citep{charbonneau}. This algorithm performs the maximization of a user defined function, for which purpose we used the inverse $\\chi^2$. Since there are many photometric measurements in the mid-infrared domain, but just a few in the far-infrared, the mid-infrared region has a higher weight during the fit, compared to the far-infrared domain. Therefore, in order to ensure an equally good fit at all wavelengths, we divided the SED into four regions and weighted the $\\chi^2$ of each domain with the inverse of number of photmetric points the region contained. Then the final $\\chi^2$ was the sum of the $\\chi^2$ of all regions. The regions we used were: $0.3-3$, $3-30$, $30-300$ and $300-3000\\,\\mu$m. The parameters of the best-fit model, which gives a weighted $\\chi^2$ of 0.67, are listed in Table~\\ref{tab:par}. The fitted model SED as well as the disc and envelope components are overplotted in Fig.~\\ref{fig:sed}. The model SED is consistent with the observed fluxes. This shows that the picture of a thin accretion disc and an envelope is consistent with both the measured SED and the geometry and disc/envelope parameters inferred from our polarimetric observations. Detailed modelling of the silicate, PAH, and ice spectral features, as well as the correct treatment of internal extinction and scattering would require radiative transfer modelling, which will be the topic of a subsequent paper. \\subsection{The evolutionary status of \\pars{}} \\label{sec:general} The geometry of FUor models discussed in the literature (e.g.~\\citealt{hk96, tbb}) usually consist of a central star surrounded by an accretion disc and an infalling envelope with a wind-driven polar hole. These assumptions are supported by the fact that they fit well the SED \\citep{green, quanz}, the interferometric visibilities \\citep{mg,v1647ori_midi} and the temporal evolution of the SED \\citep{fuors}. In this paper we present the first direct imaging of these circumstellar structures in a FUor. Our polarimetric measurements of \\pars{} show the existence of a circumstellar disc which extends from at least 48 to 360 AU. The most striking feature of the disc is its flatness over the whole observed range. The short-wavelength part of the SED could be well reproduced using a radial temperature profile of $r^{-0.75}$ (Sect.~\\ref{sec:modelling}). This profile is expected from both a geometrically thin accretion disc and a flat reprocessing disc. An envelope was also seen in the polarization maps of \\pars{} and it was also a necessary component for the SED modelling. Envelopes are involved in many FUor models and in this paper we present a direct detection of this model component. Our images reveal that the envelope can be followed inwards as close to the star as the disc. FUor models often assume a polar cavity in the envelope, created by a strong outflow or disc wind. The direct images of \\pars{} clearly show the presence of such a cavity and we also detected a bipolar outflow in the \\pars{} system. In the recent years, as new interferometric and infrared spectroscopic observations were published for FUors, the group turned out to be more inhomogeneous in physical properties than earlier assumed, when mainly optical photometry and spectroscopy had been available. \\citet{quanz} proposed that some differences might be understood as an evolutionary sequence. They suggest that FUors constitute the link between embedded Class I objects and the more evolved Class II objects. Members of the group exhibiting silicate absorption at $10\\,\\mu$m are younger and more embedded (Category 1, e.g.~V346\\,Nor); while objects with pure silicate emission are more evolved (Category 2, e.g.~FU\\,Ori and Bran\\,76). There are objects showing a superposition of silicate absorption and emission, which are probably in an intermediary evolutionary stage (e.g.~RNO\\,1B). \\citet{green} also sorted FUors, based on the ratio of the far-infrared excess and the luminosity of the central accretion disc, $f_d$ (Equ.~7 in their paper). A large relative excess ($f_d\\,{>}\\,5\\%$) indicates an envelope of large covering fraction (V1057\\,Cyg and V1515\\,Cyg), while low relative excess means a tenuous or completely missing envelope (Bran\\,76 and FU\\,Ori). This is also an evolutionary sequence, as young, more embedded objects have large envelopes, while around more evolved stars, the envelope has already dispersed. $f_d$ can also be used to calculate the opening angle of the envelope, thus a prediction of this scheme is that the opening angle is becoming wider during the evolution, probably due to strong outflows during the repeated FUor outbursts. The two classification schemes are not inconsistent and one can merge them into the following evolutionary sequence: (1) the {\\it youngest objects} exhibit silicate absorption and large far-infrared excess (V346\\,Nor, probably also OO\\,Ser and L1551\\,IRS\\,5 belong here); (2) {\\it intermediate-aged objects}, where the silicate feature is already in emission but there is still a significant far-infrared excess (V1057\\,Cyg, V1515\\,Cyg, probably also RNO\\,1B and V1647\\,Ori); (3) the most {\\it evolved objects} show pure silicate emission and low far-infrared excess (FU\\,Ori, Bran 76). We note, however, that this classification has some weak points. As \\citet{quanz} already mentioned, an edge-on geometry in a more evolved system may appear as a younger one. Moreover, during an outburst and the subsequent fading phase, certain spectral features as well as the global shape of the SED may change. \\pars{} can be placed in this evolutionary scheme, though one should keep in mind that because of the nearly edge-on geometry, the classification of this object is somewhat uncertain. \\pars{} displays silicate emission (Fig.~\\ref{fig:pah}). Integrating the flux of the two components in our simple model (Sect.~\\ref{sec:modelling}), and correcting the apparent disc luminosity for inclination effect ($i=86^{\\circ}$, Table~\\ref{tab:par}) using Equ.~6 of \\citet{green}, we obtained a large relative far infrared excess of $f_d\\,{=}\\,75\\%$. These two properties place \\pars{} into the intermediate-aged category (though because of its inclination, it may actually seem younger than it is). Following Equ.~7 of \\citet{green}, from the $f_d$ value, we also computed the opening angle of the envelope. The resulting opening angle of $60^{\\circ}$ agrees well with the angle measured in the direct NACO images (Sect.~\\ref{sec:env}). \\pars{} was placed into the evolutionary scheme using two parameters: the silicate feature and the relative far infrared excess. In the following we discuss whether its other physical characteristics match with those of other FUors. \\begin{itemize} \\item[{\\it (i)}] Our observations revealed that the circumstellar disc of \\pars{} is very flat. Due to lack of similar direct measurements for other FUors, we can only speculate that perhaps all FUors with envelopes have such flat discs. On the other hand, the most evolved FUor, FU\\,Ori, seems to have no envelope but its disc is probably flared \\citep{kh91, green, quanz2}. This might suggest that disc flaring develops at later stages, when illumination from the central source may heat the disc surface more directly. \\item[{\\it (ii)}] In a nearly edge-on system like \\pars{}, one expects to see the $10\\,\\mu$m silicate feature in absorption. The fact that \\pars{} has silicate emission indicates that the line of sight towards the central region is not completely obscured. Using the optical depth of the $15.2\\,\\mu$m CO$_2$ ice feature, we calculated an $A_V\\,{=}\\,8\\,$mag ($A_V\\,{=}\\, 38.7 A_{15.2 \\mu\\rm{}m}$, \\citealt{savage}). This value is surprisingly low compared to V1057\\,Cyg ($A_V\\,{\\sim}\\,50-100\\,$mag, \\citealt{kh91}). This indicates a much more tenuous envelope, which is also supported by the low envelope mass of $0.02\\,\\rm{}M_{\\odot}$ in our modelling. \\item[{\\it (iii)}] Following \\citet{quanz}, we analysed the profile of the $15.2\\,\\mu$m CO$_2$ ice feature of \\pars{}. The inset in Fig.~\\ref{fig:pah} shows that the feature has a characteristic double-peaked sub-structure, very similar to HH\\,46\\,IRS, an embedded young source \\citep{boogert}. HH\\,46\\,IRS is a reference case for processed ice. The presence of processed ice in \\pars{} indicates heating processes and the segregation of CO$_2$ and H$_2$O ice, already at this evolutionary stage. Other FUors exhibiting this kind of profile are L1551\\,IRS\\,5, RNO\\,1B and RNO\\,1C \\citep{quanz}. \\end{itemize} The evolutionary state of a young stellar object can also be estimated following the method proposed by \\citet{chen}. According to their Equ.~(1) we calculated a bolometric temperature of T$_{\\rm bol}=410\\,$K for the measured SED. We compared this value with the distribution of corresponding values among young stellar objects in the Taurus and $\\rho\\,$Ophiuchus star forming regions (Chen et al. 1995). From this check we can conclude that \\pars{} seems to be a class I object, and its age is ${\\sim}\\,10^5\\,$yr. However, \\citet{green} argued that the apparent SED of the disc component depends on the inclination. Thus we computed T$_{\\rm bol}$ also for a face-on disc configuration and obtained T$_{\\rm bol}=1160\\,$K, corresponding to a Class II object. In fact, \\pars{} is probably close to the Class I / Class II border, in accordance with the proposal of \\citet{quanz}." }, "0710/0710.3328_arXiv.txt": { "abstract": "We analyze the electromotive force (EMF) terms and basic assumptions of the linear and nonlinear dynamo theories in our three-dimensional (3D) numerical model of the Parker instability with cosmic rays and shear in a galactic disk. We also apply the well known prescriptions of the EMF obtained by the nonlinear dynamo theory (Blackman \\& Field 2002 and Kleeorin et al. 2003) to check if the EMF reconstructed from their prescriptions corresponds to the EMF obtained directly from our numerical models. We show that our modeled EMF is fully nonlinear and it is not possible to apply any of the considered nonlinear dynamo approximations due to the fact that the conditions for the scale separation are not fulfilled. ", "introduction": "\\label{sec:intro} It seems that the issue of the magnetic field amplification in galaxies may be well explained by the two main physical mechanisms: the Parker instability (PI), which takes into account the cosmic rays (CR) and the shear \\cite[e.g.][and references therein]{hanasz03,hanasz04}, and the magneto-rotational instability \\cite[MRI, e.g.][]{dziourkevitch04,kitchatinov04}. The possible scenario of the magnetic field evolution could be presented as follows: when the protogalaxy starts to rotate differentially, the MRI mechanism occurs, and this results in a very efficient magnetic field amplification even to the level of $\\mu$G with the global e-folding time of 100~Myr or even less \\citep{dziourkevitch04}. Simultaneously, the quadrupole symmetry of the large-scale magnetic field is being created. MRI also causes the turbulent motions in the galactic disks. The process of supernovae (SN) explosions, that arises in young galactic objects \\cite[e.g.][and references therein]{widrow02} suppresses the MRI mechanism. In the same time, the cosmic rays produced in SN remnants may induce the Parker instability process \\cite[e.g.][]{parker92,hanasz04}. Hence, we may conclude that during the early stage of galaxy evolution the MRI process is being replaced by the PI mechanism. The local simulations of the large-scale magnetic field took into account the turbulent dynamo theory and the magnetic back-reaction onto the turbulent motions \\citep{piddington70,piddington72a,piddington75a,kulsrud95}. The authors drew the conclusion that it was difficult to obtain the amplification of the total magnetic field \\cite[e.g.][and references therein]{widrow02}. It might be explained by the equipartition of the random magnetic field component with the random turbulent motions. Such process suppresses the dynamo action at later times. Moreover, the magnetic field amplification might be easily stopped even by the presence of the weak large-scale magnetic field \\citep{cattaneo91,vainshtein92,cattaneo94,cattaneo96,ziegler96}. \\cite{blackman99} explained that papers analyzing the analytically strong suppression of the dynamo coefficient $\\alpha$ should distinguish between different state orders of the turbulent quantities. However, their analysis did not completely solve the problem of the quenching of the dynamo coefficients. In their next paper \\citep{blackman00}, they proved that the results from the \\cite{cattaneo96} model were based on the assumption about the periodicity of the boundary conditions. Nevertheless, the quenching effects could also appear even when the open boundary conditions were applied \\cite[e.g.][]{brandenburg01b}. The classical dynamo theory does not conserve the total magnetic helicity \\cite[see e.g.][BF02]{blackman02}. In media characterized by the high magnetic Reynolds number ($R_m\\gg1$) the total helicity should remain constant in closed regions \\cite[e.g.][]{berger84,brandenburg02b,brandenburg05b,subramanian02}. The permanent helicity is also an additional factor, which suppresses the dynamo activity \\citep[e.g.][]{brandenburg02b}. The following papers \\citep{blackman00,kleeorin00,blackman03,kleeorin99,kleeorin02,kleeorin00, kleeorin03,rogachevskii00,rogachevskii01} presented the two methods that allowed the modeling of the dynamo action evading the problem of the constant helicity. The first method is based on the ejection of the magnetic helicity through boundaries. The second one uses the creation of the negative and positive helicity at the large and small scales respectively \\cite[see][]{brandenburg02b,kleeorin02,kleeorin03}. \\cite{blackman02} in their next paper analyzed the nonlinear prescription of both dynamo coefficient $\\alpha$ and $\\beta$. Both factors were obtained without any linearization and took into account all terms in the equation of the evolution of the fluctuating part of the magnetic field (see Eq.~\\ref{eqn:fluct_field_evol}). The results were similar to the dynamo coefficients obtained by \\citep{pouquet76}, but the units were different (without the time integration). They also solved the small- and large-scale helicity dynamo equations numerically simultaneously with the equation for the EMF time evolution. That allowed them to obtain the growth of the large-scale magnetic field in the kinematic phase. The new form of the dynamo coefficients for anisotropic turbulent motions with the presence of the large-scale magnetic field was presented by \\cite{rogachevskii01}. They calculated dynamo coefficients according to the \\cite{raedler80} EMF prescription \\cite[see also][]{kowal05}, which neglected all quadratic terms in the mean field in the EMF. \\cite{kleeorin03} used those forms of the dynamo coefficients to solve numerically the dynamo equation in the local thin-disc approximation. The authors took into account the quenching of both coefficients, $\\alpha$ and $\\beta$, and helicity flux through the boundary. They found that it was possible to obtain the growth of the large-scale magnetic field when $\\alpha$-quenching was only analyzed \\citep{kleeorin02}. If the model included also $\\beta$-quenching, no growing solution of the dynamo \\citep{kleeorin03} could be obtained. Both in Blackman-Field and in Kleeorin-Rogachevskii approaches there is an $\\alpha_{\\rm m}$ term that quantifies the small scale helicity current. This term depends on the current helicity flux, and there are different theories for this flux. In Kleeorin et al. a heuristically motivated expression for the flux was used, in \\cite{brandenburg05b} the \\cite{vishniac01} flux was used, and in \\cite{shukurov06} a simple advective flux was used. In all these cases the current helicity flux allows the field to saturate at high levels. The latest research on the turbulent enforcement in the solar convective zone calculated in the local cube with the shear has shown that the open boundaries help to obtain the amplification of the large-scale magnetic field even without the helicity of turbulent motions \\cite[end references therein]{brandenburg05a,brandenburg05b}. We have to stress that their result, an increase of the total magnetic energy, was obtained without any assumption considering the additional EMF of dynamo. The authors applied isotropic and homogeneous turbulence with and without the helical forcing in their model. On the other hand, Brandenburg and his collaborators \\citep{brandenburg05a} interpreted their results in terms of the mean field dynamo theory. The time evolution of $\\alpha$ was obtained from the calculated electromotive force \\citep{brandenburg04,kowal05,brandenburg02a,brandenburg01a}. \\cite{brandenburg04} explained that thanks to the flux of the current helicity flowing out of the cube the process of $\\alpha$-quenching tends to be not as disastrous as forseen. The only conditions are the intermediate level of $R_m$ and open boundaries. When the high value of $R_m$ is applied to the model, the comparatively lower value of the large-scale magnetic field strength are obtained \\citep{brandenburg04}. However, in astrophysical objects $R_m$ is always high. That is why this result seems to be peculiar. On the other hand, the authors made it clear that the total magnetic energy grows mainly in the kinematic phase of the dynamo and their results do not depend strongly on $R_m$. Futhermore, they calculated the value of $\\alpha$ coefficient based on the modeled EMF. The obtained factor was similar to the same coefficient calculated according to the \\cite{kleeorin00,kleeorin02,kleeorin03} prescriptions. We believe that the fact that such similarity occurs results from the isotropic and homogeneous turbulence in both models \\citep{brandenburg05a,brandenburg05b,brandenburg04}. It may also happen due to the fact that the authors applied standard dynamo approximations, which neglect the quadratic terms in the mean field in EMF. The previously mentioned results indicate that the realistic physical simulations are of the great importance when the MRI \\citep{dziourkevitch04} and the PI \\citep{hanasz04,hanasz03,kowal05} processes are considered. Both models, which meet enumerated requirements, showed that it was possible to amplify galactic magnetic field efficiently. \\cite{hanasz03} and \\cite{hanasz04,hanasz05} presented that the following two processes: the Parker instability driven by cosmic rays from supernovae and the shear from the differential rotation enable the magnetic field amplification (with the e-folding time scale of 250~Myr or even 140~Myr). The model also applied realistic gravity according to the \\cite{ferriere98} prescriptions. The idea of obtaining the dynamo coefficients ($\\alpha$ and $\\beta$) from the electromotive force calculated in the local numerical simulations proved to be essential for many other authors too \\cite[e.g.][etc]{ziegler96,brandenburg02a}. We included the calculations of the dynamo coefficients, which we obtained from the calculated EMF in our previous paper \\citep{kowal05}. The application of the statistical methods provided us with the acceptable values of the dynamo $\\alpha$-tensor. On the other hand, the values of $\\beta$ coefficients were negative. Such values are inconsistent with the R\\\"adler prescription \\citep{raedler80}. This may be caused either by the applied statistical method, which does not take into account physical differentiation, or by the linear EMF approximation \\citep{kowal05}. For this reason we decided to analyze that matter in our present study. We search for the conditions, which should be fullfiled in order to make linearization of the electromotive force in the mean field dynamo theory possible. We would like to examine the following problems: the scale separation, the ratios of the terms in the equation for the small-scale magnetic field evolution \\cite[e.g.][]{raedler80}, the magnitude of the turbulent kinetic energy in comparison to the large-scale magnetic one. Next, we plan to apply the estimations of the dynamo coefficients derived by \\cite{blackman00,rogachevskii01} and \\cite{brandenburg04} to our models. We would like to check if their approximations fit into the calculated electromotive force in our models \\cite{hanasz04}. In this part of this work we do not include the explicit analysis involving the conservation of the magnetic helicity. The investigation of the magnetic helicity conservation in our models is already advanced, but it is complex enough to be described in separate paper, which is under preparation. We cannot apply the approximations of \\cite{ruediger93} and \\cite{kitchatinov94} derived from the EMF quenching by the usage of the Second Order Correlation Approximations. They assumed that the Strouhal number $S$ is essentially smaller than 1 ($S\\ll1$, where $S = \\tau_c \\times \\rm v/ \\rm l_c$). This assumption is not fullfiled in our numerical experiments, where $S$ is about 1. Finally, we discuss our results. \\begin{table*} \\begin{center} \\caption{Parameters of models examined in this paper. (*) The conversion rate, presented as 10\\% in \\cite{hanasz04}, was in fact equal to 100\\%, due to a trivial calculation mistake. Therefore, the overall injection rate of cosmic ray energy was equivalent to a realistic one, corresponding to SN rate=~20~kpc$^{-2}$Myr$^{-1}$, with the energy conversion factor = 10\\%. \\label{table-params}} \\begin{tabular}{|l|c|c|c|c|} \\hline Model & A & B & C & D\\\\ \\hline Domain sizes [kpc] & 0.5 $\\times$ 1 $\\times$ 1.2 & \\multicolumn{3}{c}{0.5 $\\times$ 1 $\\times$ 4} \\vline \\\\ Resolution & 50 $\\times$ 100 $\\times$ 120& \\multicolumn{3}{c}{50 $\\times$ 100 $\\times$ 400} \\vline \\\\ Vertical gravity at $R$ [kpc] = & 8.5 & \\multicolumn{3}{c}{5} \\vline\\\\ Gas column density [cm$^{-2}$] & ... & \\multicolumn{3}{c}{27$\\times$10$^{20}$} \\vline\\\\ Angular velocity $\\Omega$ [Myr$^{-1}$] & 0.05 & \\multicolumn{3}{c}{0.05} \\vline \\\\ SN rate [kpc$^{-2}$Myr$^{-1}$] & $2$ & \\multicolumn{3}{c}{130} \\vline \\\\ Initial $\\alpha=e_{\\rm mag}/e_{\\rm gas}$ & $10^{-8}$ & \\multicolumn{3}{c}{$10^{-4}$}\\vline \\\\ Diffusion coefficients $K_\\parallel$, $K_\\perp$ [cm$^2$s$^{-1}$] & $3 \\times 10^{27}$, $3 \\times 10^{26}$ & \\multicolumn{3}{c}{$3 \\times 10^{27}$, $3 \\times 10^{26}$}\\vline \\\\ Conversion rate of SN kinetic to CR energy & 10\\%$^{(*)}$ & \\multicolumn{3}{c}{10\\%}\\vline \\\\ \\hline Resistivity $\\eta$ [cm$^2$s$^{-1}$] & $3\\times 10^{24}$ & $0\\times 10^{24}$ & $3\\times 10^{24}$ & $30\\times 10^{24}$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} ", "conclusions": " \\begin{enumerate} \\item {Neither the velocity nor the magnetic field scale separation occurs in our model.} \\item { The electromotive forces in the cosmic-ray driven dynamo model are nonlinear, but none of the two examined nonlinear approaches is capable of reproducing electromotive forces in the numerical experiments correctly.} \\item {Various nonlinear prescriptions of the dynamo coefficients have been proposed by other outhors, however, they are not capable of reconstructing the electromotive force resulting from experiments of cosmic-ray driven dynamo. Moreover the reconstructions of the magnetic field produce too fast or too slow growth of the magnetic energy in comparison with the results of cosmic-ray driven dynamo numerical experiments.} \\end{enumerate} Extension of the present work, including considerations of magnetic helicity conservation will be presented in the forthcoming paper." }, "0710/0710.1094_arXiv.txt": { "abstract": "We report a measurement of the supernova (SN) rates (Ia and core-collapse) in galaxy clusters based on the 136 SNe of the sample described in \\citet{C99} and \\citet{M05}. Early-type cluster galaxies show a type Ia SN rate (0.066 SNuM) similar to that obtained by \\citet{sharon07} and more than 3 times larger than that in field early-type galaxies (0.019 SNuM). This difference has a 98\\% statistical confidence level. We examine many possible observational biases which could affect the rate determination, and conclude that none of them is likely to significantly alter the results. We investigate how the rate is related to several properties of the parent galaxies, and find that cluster membership, morphology and radio power all affect the SN rate, while galaxy mass has no measurable effect. The increased rate may be due to galaxy interactions in clusters, inducing either the formation of young stars or a different evolution of the progenitor binary systems. We present the first measurement of the core-collapse SN rate in cluster late-type galaxies, which turns out to be comparable to the rate in field galaxies. This suggests that no large systematic difference in the initial mass function exists between the two environments. ", "introduction": "\\label{sec:intro} Type Ia Supernovae (SNe Ia) are believed to be the result of the thermonuclear explosion of a C/O white dwarf (WD) in a binary system due to mass exchange with the secondary star. This conclusion follows from a few fundamental arguments: the explosion requires a degenerate system, such as a white dwarf; the presence of SNe Ia in old stellar systems implies that at least some of their progenitors must come from old, low-mass stars; the lack of hydrogen in the SN spectra requires that the progenitor has lost its outer envelope; and, the released energy per unit mass is of the order of the energy output of the thermonuclear conversion of carbon or oxygen into iron. Considerable uncertainties about the explosion model remain within this broad framework, such as the structure and the composition of the exploding WD (He, C/O, or O/Ne), the mass at explosion (at, below, or above the Chandrasekhar mass) and the flame propagation (detonation, deflagration, or a combination of the two). The key observations constraining the explosion models are the light curve and the evolution of the spectra. Large uncertainties also remain regarding the nature of the progenitor binary system, its evolution through one or more common envelope phases, and its configuration (single or double-degenerate) at the moment of the explosion (see \\citealt{yungelson05}, for a review). Solving the problem of the progenitor system is of great importance for modern cosmology as SNe dominate metal production, (e.g., \\citealt{matteucci86}), are expected to be important producer of high-redshift dust \\citep{maiolino01, maiolino04a,maiolino04b,bianchi07}, and are essential to understand the feedback process during galaxy formation (e.g., \\citealt{scannapieco06}). The nature of the progenitor systems can be probed by studying the SN rate in different stellar populations, and constraining the delay time distribution (DTD) between star formation and SN explosion. \\smallskip In 1983, Greggio \\& Renzini computed the expected DTD for a single-degenerate system. The computation was later refined by many authors and extended to double-degenerate systems \\citep{tornambe86,tornambe89,tutukov94,yungelson00,matteucci01,belczynski05,greggio05}. The DTD can be convolved with the star formation history (SFH) of each galaxy to obtain its SN rate. The observation of the SN rates per unit mass in galaxies of different types \\citep{M05,sullivan06} and in radio-loud early-type galaxies \\citep{dellavalle05} has proved to be an effective way to constrain the DTD. The SN rates per unit mass show that SNe Ia must come from both young and old progenitors \\citep{M05,sullivan06}. The dependence of the SN rate on the radio power of the parent galaxy is well reproduced by a ``two channel'' model \\citep{mannucci06}, in which about half of the SNe Ia, the so-called ``prompt'' population, explode soon after the formation of the progenitors, on time scales shorter than $10^8$ yr, while the other half (the ``tardy'' population) explode on a much longer time scale, of the order of $10^{10}$ yr. Several attempts to compare the evolution of SN rate with redshift with that of the SFR have also been presented (see, among many others, \\citealt{galyam04,dahlen04,cappellaro05,neill06,barris06,botticella07} and \\citealt{poznanski07}), but the large uncertainties on both quantities prevent strong conclusions (see, for example, \\citealt{forster06}). \\smallskip In principle, an accurate measurement of the DTD could identify the progenitor binary system. In practice, both the large number of free parameters involved in the theoretical computations of the DTD, and the complex SFHs of most of the galaxies make this identification much more uncertain. To solve the problem of the complexity of the SFH, it is interesting to measure the SN Ia rate in galaxy clusters. Most of the stellar mass of these systems is contained in elliptical galaxies, whose stellar populations are dominated by old stars. Despite the problem that even a small amount of new stars could give a significant contribution to the SN rate (see the discussion in sect.~\\ref{sec:discussion}), the reduction in the uncertainty in the SFH is of great help to derive the DTD. \\smallskip The cluster SN rate is also of great importance to study the metallicity evolution of the universe. The gravitational potential well of galaxy clusters is deep enough to retain in the intracluster medium (ICM) all the metals which are produced in galactic or intergalactic SNe. As a result, the metallicity of the ICM is a good measure of the integrated past history of cluster star formation and metal production. As discussed by \\citet{renzini93}, the measured amount of iron is an order of magnitude too high to be produced by SNe Ia exploding at the current rate. Explanations of this effect include the presence of higher SN rates in the past \\citep{matteucci06}, the importance of the intracluster stellar population \\citep{zaritsky04}, or evolving properties of star formation processes \\citep{maoz04,lowenstein06}. The observed abundance ratios in the ICM can be used to constrain the ratio between the total numbers of Ia and CC SNe, as recently done by \\citet{deplaa07}. Constraints on the SN Ia models can also be derived from the radial distribution of metallicity \\citep{dupke02}. \\citet{calura07} used the observed cosmic evolution of iron abundances in \\citet{balestra07} to constrain the history of SN explosion, iron formation and gas stripping in galaxy clusters. They found good agreement with the observations, especially when the ``two channel'' model of SNe Ia by \\citet{mannucci06} is used. \\smallskip There are strong motivations for measuring also the cluster rates of the other physical class of SNe, the core-collapse (CC) group. Type II and type Ib/c SNe are attributed to this group because there is a general consensus that these explosions are due to the collapse of the core of a massive (about 8--40~\\msun) star. Thus, CC SNe are expected to be good tracers of star formation in moderately dusty environments (see \\citealt{mannucci07}). Their rate per unit mass is also very sensitive to the initial mass function (IMF), because SN explosions are due to massive stars while most of the mass is locked in low-mass stars. As a consequence, studying the CC SN rate as a function of environment is a sensitive test for any systematic difference in IMF. \\begin{table} \\caption{ Measured type Ia SN rates in early-type cluster galaxies \\label{tab:history} } \\begin{tabular}{llcc} \\hline \\hline Reference & ~~$z$ ~~($z$ range) & N$_{SN}$& Rate\\\\ & & & (SNuB) \\\\ \\hline This work & 0.02 (0.005--0.04) & 20 & $0.28^{+0.11}_{-0.08}$\\\\ \\citet{crane77} & 0.023 (0.020--0.026)& 8 & $\\sim$0.10 \\\\ \\citet{barbon78} & 0.023 (0.020--0.026)& 5 & $\\sim$0.16 \\\\ \\citet{germany04} & 0.05 (0.02--0.08) & 23 & unpubl. \\\\ \\citet{sharon07} & 0.15 (0.06--0.19) & 6 & 0.27$^{+0.16}_{-0.11}$\\\\ \\citet{galyam02} & 0.25 (0.18--0.37) & 1 & 0.39$^{+1.65}_{-0.37}$\\\\ \\citet{galyam02} & 0.90 (0.83--1.27) & 1 & 0.80$^{+0.92}_{-0.41}$\\\\ \\hline \\end{tabular} \\end{table} \\subsection{The observed cluster supernova rate} Prompted by all these motivations, several groups have measured the SN Ia rate in galaxy clusters, but the results are still quite sparse. The first published values are due to \\citet{crane77} and \\citet{barbon78} (see Table~\\ref{tab:history} for a summary, including the results of our work, discussed below), before a clear distinction between type Ia and Ib/c had been introduced. They used a sample of 5--8 SNe in the Coma Cluster and constrained the SN rate to be of the order of 0.15 SNuB (SN per century per $10^{10}$ \\lsun\\ in the B band). The SN rate as a function of galaxy environment was also addressed by \\citet{caldwell81} to derive information on SN progenitors. Modern searches for cluster SNe begin with \\citet{norgaard89} who discovered a SN Ia in a cluster at $z=0.31$. Starting from the late '90s, the Mount Stromlo 1.3 m telescope was used to monitor a few tens of Abell Clusters \\citep{reiss98}. Three years of monitoring resulted in the detection of 23 candidate SNe Ia in cluster galaxies \\citep{germany04}, but a rate based on this sample was never published. The first rates for cluster galaxies based on modern searches were published by \\citet{galyam02}. These authors used archive images from the Hubble Space Telescope (HST) of 9 galaxy clusters, and discovered 6 SNe, 2 of which are associated with the clusters, at $z=0.18$ and $z=0.83$. The derived rates were affected by large statistical uncertainties due to the small number of detected SNe, but were consistent with a moderate increase of the rate with redshift compared to the rate in local elliptical galaxies. A sample of 140 low-redshift Abell clusters were monitored by the Wise Observatory Optical Transient Search (WOOTS, \\citealt{galyam07}) using the Wise 1m telescope. The seven detected cluster SNe were used to constrain the fraction of intergalactic stars and SNe \\citep{galyam03} and to measure the cluster SN rate \\citep{sharon07}. This latter work obtains a value of the SN rate per unit mass of $0.098^{+0.058}_{-0.039}$ SNuM (SN per century per $10^{10}$ \\msun\\ of stellar mass), which is larger than, but still consistent with, the value of $0.038^{+0.014}_{-0.012}$ SNuM, derived by \\citet{M05} for local ellipticals. Finally, a SN search in clusters is ongoing at the Bok Telescope on Kitt Peak \\citep{sand07}. All of the previous published SN Ia rates are based on a small number of SNe and, as a consequence, have large statistical errors. Also, a cluster rate for CC SNe has never been published because many of the cited samples only contain Ia SNe. In this work, we use the SNe in the \\citet{C99} sample to study the SN rate as a function of galaxy environment. Throughout this paper we use the ``737'' values of the cosmological parameters: $(h_{100},\\Omega_m,\\Omega_\\Lambda)=(0.7,0.3,0.7)$. \\begin{figure} \\includegraphics[width=9cm]{galdist.ps} \\caption{ \\label{fig:galdist} Surface density of galaxies (upper panel) and fraction of early-type galaxies (lower panel) as a function of the projected distance from the closest cluster. Above 3 Mpc, the average for field galaxies is shown. The vertical dotted lines show the two projected distances, 0.5 and 1.5 Mpc, used to define cluster galaxies (see text). } \\end{figure} ", "conclusions": "\\label{sec:discussion} The interpretation of the possible difference in SN Ia rate between cluster and field early-type galaxies is not straightforward. As the observed rate is the convolution of the SFH with the DTD, the differences could be due to either of these functions. \\begin{enumerate} \\item The first possibility is that the rate difference is due to differences in the stellar populations. \\citet{M05}, \\citet{sullivan06}, and \\citet{aubourg07} have shown that the type Ia SN rate has a strong dependence on the parent stellar population, with younger stars producing more SNe. The difference in SN rate could be related to this effect, i.e., to a higher level of recent star formation in cluster ellipticals. Only a very small amount of younger stars is needed, because the amplitude of the DTD at short times can be hundreds of times larger than at long times. As an example, the \\citet{greggio83} single-degenerate model has 300 times more amplitude at $10^8$ yr than at $10^{10}$ yr, and this means that a recently formed stellar population contributing 0.3\\% of the mass can provide as many SNe as the remaining 99.7\\% of old stars. For the ``two channel'' model by \\citet{mannucci06}, the amount of young stars needed can be even lower, at the 0.1\\% level, as this DTD amplitude ratio between $10^7$ and $10^{10}$ years is as large as 1000. \\smallskip The presence of traces of star formation in early-type galaxies is not inconsistent with other observations. Many ellipticals show signs of recent interactions or star formation activity: faint emission lines \\citep{sarzi06}, tidal tails \\citep{vandokkum05}, dust lanes \\citep{vandokkum95,colbert01}, HI gas \\citep{morganti06}, molecular gas \\citep{welch03}, and very blue UV colors \\citep{kaviraj06,schawinski07,haines07,kaviraj07}. Even if the interpretation of most of these effects is matter of debate (for example, \\citealt{serego07} have found only small amounts of HI gas in cluster ellipticals), the observations suggest a widespread, low-level presence of star formation. The dependence of this presence with environment is not settled yet. \\citet{ferreras06} have found evidence for recent star formation, at the percent level, in ellipticals in compact groups, but not in field ellipticals. In contrast, \\citet{verdugo07} and \\citet{haines07} have found higher levels of present star formation in field rather then cluster early-type galaxies. \\smallskip Some studies (see, for example, \\citealt{sanchez06}, \\citealt{bernardi06} and \\citealt{collobert06}), have found younger ages in field early-type galaxies with respect to cluster galaxies (but \\citealt{serego06} have found no difference). Taken at face value, this would seem to contradict the star formation interpretation of the SN rate, but this is not necessary the case. Field ellipticals could be younger that cluster ellipticals, but nevertheless they could show a lower level of {\\em present} star formation. The difference in the age of the {\\em dominant} stellar population of early-type galaxies, of the order of 1 Gyr for ages of about 12 Gyr, might not be directly related to the amount of star formation in the last few $10^8$ years. Such a contribution cannot be detected in the integrated colors of the galaxies. The expected differences are at the 0.05 mag level for the (B--K) color, assuming the younger stars are not associated with dust, and even smaller (0.02 mag for $A_V$=1), allowing for dust extinction. It is usually assumed that early-type galaxies can form new stars only after merging with a small, gas rich galaxy, because usually they do not host much interstellar gas. The average amount of stars formed is proportional to the merger (or encounter) rate, to the typical amount of gas in the accreted galaxy, and to the efficiency of star formation in the accreted gas. It is possible that one or more of these quantities are larger for cluster galaxies than for field galaxies because of the different galaxy volume density and galaxy-galaxy encounter velocity. \\smallskip If this is the correct interpretation, the ``prompt'' population of SNe Ia would be associated with the explosion of CC SNe from the same young stellar populations. If a SN Ia is to explode within $10^8$ yr of the formation of its progenitor, the primary star of the progenitor binary system must have a mass above 5.5 \\msun\\ to allow for the formation of a white dwarf in such a short time. \\citet{mannucci06} have shown that reproducing the observed SN rates by using the ``bimodal'' DTD in that paper implies that about 7\\% of all stars between 5.5 and 8 \\msun\\ explode as ``prompt'' SNe Ia, while the ``tardy'' population corresponds to a lower explosion efficiency, about 2\\%, and on a much longer timescale (see also \\citealt{maoz07} for various estimates of these efficiencies). For a Salpeter IMF and assuming that 100\\% of the stars between 8 and 40 \\msun\\ end up at CC SNe, we expect 1.3 CC SNe for each ``prompt'' type Ia. Assuming that the difference between cluster and field early-type galaxies is due to the ``prompt'' SNe Ia, the rate of this population is of the order of 0.066-0.019=0.047 SNuM (see Table~\\ref{tab:massrate}). Converting this rate to an observed number, about 2 CC SNe are expected in the cluster early-type galaxies of our sample, consistent with our null detection at about 1.3$\\sigma$ level. We conclude that the non detection of CC in the early-type galaxies belonging to our sample and the corresponding upper limits to the CC rate are consistent with the hypothesis of a ``prompt'' Ia component. We also note that some CC SNe have been discovered in the recent past in prototypical early-type galaxies. \\citep{pastorello07}. \\item A second possible interpretation is that the higher rate in cluster early-type galaxies is related to differences in the DTD. If the stars in ellipticals are 9-12 Gyrs old (see, for example, \\citealt{mannucci01}), the SN rate is dominated by the tail of the DTD at long times. Differences in the environments could produce small differences in the shape of this function, for example because of the higher numbers of encounters. A interesting possibility is also that the changes in the DTD are related to differences in metallicity between cluster and field early-type galaxies, as discussed by \\cite{sanchez06,bernardi06,collobert06} and \\cite{prieto07}. The differences between cluster and field galaxy metallicity presented by these papers are neither large nor always in the same direction. Nevertheless systematic, although not large, differences in metallicity could be present and produce significant changes in the DTD, for example, by affecting the efficiency of mass loss during the complex life of a binary system. \\end{enumerate} Table~\\ref{tab:history} lists the different measurements of the SN rate in early-type cluster galaxies. The evolution of this rate can be compared with the history of star formation of the parent galaxies to derive the DTD. Currently published cluster SN rates at z$>$0.2 are too uncertain to permit any strong conclusions. However, current and future searches for SNe are expected to change this situation and allow for the derivation of meaningful constraints (see, for example, \\citealt{sharon06}). \\smallskip To summarise, we have used a sample of 136 SNe in the local universe to measure the SN rate as a function of environment. For the first time, we measure the CC SN rate in clusters. We find it is very similar to the CC SN rate in field galaxies, suggesting that the IMF is not a strong function of the environment. For Ia SNe, the rates in clusters and in the field are similar for all galaxy types except for the early-type systems, where we detect a significant excess in clusters. This excess is not related to other properties of those galaxies, such as mass, morphology, or radio loudness. Environments itself appears to be important. We interpret this effect as possibly due to galaxy-galaxy interaction in clusters, either producing a small amount of young stars (of the order of the percent in mass over one Hubble time), or affecting the evolution of the properties of the binary systems. \\bigskip {\\bf Acknowledgments} We thank Sperello di Serego and the MEGA group (Arcetri Extragalactic Meeting) for useful discussions about the properties of elliptical galaxies. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. DM, MD, and AG thank the Kavli Institute for Theoretical Physics for its hospitality. This research was supported in part by the National Science Foundation under Grant No. PHY05-51164." }, "0710/0710.1788_arXiv.txt": { "abstract": "{The {\\it ROSAT} X-ray source \\cal\\ has recently been identified as a likely compact object whose properties suggest it could be a very nearby radio millisecond pulsar at $d = 80 - 260$\\,pc.} {We investigated this hypothesis by searching for radio pulsations using the Westerbork Synthesis Radio Telescope.} {We observed \\cal\\ at 385 and 1380\\,MHz, recording at high time and frequency resolution in order to maintain sensitivity to millisecond pulsations. These data were searched both for dispersed single pulses and using Fourier techniques sensitive to constant and orbitally modulated periodicities.} {No radio pulsations were detected in these observations, resulting in pulsed radio luminosity limits of $L_{400}^{\\rm max} \\approx 0.3 (d/250 {\\rm pc})^2$\\,mJy kpc$^2$ and $L_{1400}^{\\rm max} \\approx 0.03 (d/250 {\\rm pc})^2$\\,mJy kpc$^2$ at 400 and 1400\\,MHz respectively.} {The lack of detectable radio pulsations from \\cal\\ brings into question its identification as a nearby radio pulsar, though, because the pulsar could be beamed away from us, this hypothesis cannot be strictly ruled out.} ", "introduction": "Recently, \\citet*{rfs07}, hereafter RFS07, have identified the X-ray source \\cal\\ (from the {\\it ROSAT} All-Sky Survey Bright Source Catalog) as having an X-ray to optical flux ratio $F_{\\rm X}(0.1-2.4 {\\rm keV})/F_{\\rm V} > 8700$ ($3\\sigma$). Such a high ratio is strong evidence that \\cal, dubbed ``Calvera'' by RFS07, is a compact object. RFS07 consider several specific source classes to explain \\cal's X-ray spectrum and luminosity: an X-ray dim isolated neutron star (INS), an anomalous X-ray pulsar (AXP), a compact central object (CCO), or a nearby radio pulsar. Based on careful comparison of \\cal's properties with the canonical features displayed by these different classes of neutron star, RFS07 conclude that the most likely explanation is that \\cal\\ is a very nearby radio pulsar ($d = 80-260$\\,pc) similar to the radio millisecond pulsars (MSPs) residing in the globular cluster 47~Tuc. We have investigated this interpretation by conducting sensitive searches for radio pulsations using the Westerbork Synthesis Radio Telescope (WSRT) in the Netherlands. No pulsations were found by these searches, and we use these non-detections to place strong limits on the pulsed radio luminosity of \\cal. These limits bring into question the interpretation of this object as a nearby radio pulsar. ", "conclusions": "No plausible astronomical radio pulsations or bright, dispersed single pulses were detected in any of our observations. Using the radiometer equation modified for pulsar signals \\citep{dtws85}, we can place limits on \\cal's flux density in these observations (Table~\\ref{obs.tab}). We find that \\cal\\ has a maximum flux density at 400\\,MHz $S_{400}^{\\rm max} \\approx 4$\\,mJy and a maximum flux density at 1400\\,MHz $S_{1400}^{\\rm max} \\approx 0.3$\\,mJy, where the fractional uncertainty on these limits is roughly 50\\%. RFS07 argue that \\cal\\ may be a radio pulsar with a nearby distance $d = 80-260$\\,pc. Assuming the pulsar is isolated, we were sensitive to spin periods encompassing the observed range for radio pulsars ($P_{\\rm spin} \\sim 1$\\,ms$-10$\\,s), with a factor of roughly $2-5$ degredation in sensitivity due to red-noise at the longest periods. Using an assumed distance $d = 250$\\,pc, we can convert our flux density limits to pseudo luminosity ($L \\equiv Sd^2$) limits. We find $L_{400}^{\\rm max} \\approx 0.3 (d/250 {\\rm pc})^2$\\,mJy kpc$^2$ and $L_{1400}^{\\rm max} \\approx 0.03 (d/250 {\\rm pc})^2$\\,mJy kpc$^2$. A simple check of the ATNF pulsar catalog\\footnote{Available at http://www.atnf.csiro.au/research/pulsar/psrcat} \\citep{mhth05} shows that $\\lesssim 1$\\% of the known pulsars have a luminosity below these limits. This suggests that if \\cal\\ is a radio pulsar, then it is either especially weak, not beamed towards the Earth, or significantly further away than 250\\,pc. Assuming \\cal\\ is a radio MSP, what is the probability that its radio beam will pass the Earth (i.e. what is the beaming fraction of such pulsars)? A period-dependent beaming fraction proportional to $P_{\\rm spin}^{-0.5}$ has been shown for normal, un-recycled pulsars \\citep[see e.g.][]{ran93}. When extrapolated to millisecond spin periods, this relation implies a $\\sim 100$\\% beaming fraction for MSPs. However, \\citet{kxl+98} find observational evidence that this relationship does not apply directly to MSPs and that the beaming fraction of MSPs is more like $50-90$\\%. These estimates are consistent with considerations of the MSP population in 47~Tuc. \\citet{hge+05} used a deep {\\it Chandra} observation of the cluster to perform a population analysis which concluded that there are likely $\\sim 25$ radio MSPs ($< 60$ at 95\\% confidence) residing in the cluster, independent of beaming. Comparing this with optical and radio studies of the MSP population by \\citet{egh+03} and \\citet{mdca04} respectively, who both conclude that the total MSP population in 47~Tuc is $\\sim 30$, \\citet{hge+05} conclude that the beaming fraction is $\\gtrsim 37$\\%. While these various studies suggest that the beaming fraction of MSPs is significantly larger than for normal pulsars, and could be quite high, we cannot strongly rule out beaming as the cause of our non-detection of \\cal. It is also possible, though we feel unlikely, that we have not seen \\cal\\ because it is in a compact binary orbit and/or is highly accelerated by a binary companion. Our luminosity limits are for a coherent search and do not include these effects. However, given that we have employed acceleration searches, which are generally sensitive in cases where the total integration time $T_{\\rm int}$ is less than roughly 1/10 of the orbital period $P_{\\rm orb}$, \\cal\\ would have to be in a $\\lesssim 2$-hr orbit {\\it and} fairly weak (or eclipsed) not to have been detected in our searches of 15-min data sections. The shortest orbital period of any known radio MSP is 1.6\\,hr \\citep[][PSR~J0024$-$7204R in the GC 47~Tuc]{clf+00}. In conclusion, given the caveats we have discussed, primarily pertaining to extreme orbital or spin parameters, we believe that these searches were sensitive enough to detect any nearby radio pulsar coincident with \\cal, assuming it is beamed towards the Earth. We have checked for catalogued radio sources in the FIRST, NVSS, and WENSS surveys and find no counterpart to \\cal. Future, deeper observations detecting a radio, optical, or X-ray pulsar wind nebula could address whether this source is a nearby radio pulsar beamed away from Earth. Alternate scenarios for \\cal's nature, for instance that it might be the first unhosted CCO (RFS07), should continue to be investigated." }, "0710/0710.0351_arXiv.txt": { "abstract": "{Modern observations and models of various astrophysical objects suggest that many of their physical parameters fluctuate substantially at different spatial scales. The rich variety of the emission processes, including Transition Radiation but not limited to it, arising in such turbulent media constitutes the scope of Stochastic Theory of Radiation. We review general approaches applied in the stochastic theory of radiation and specific methods used to calculate the transition radiation produced by fast particles in the magnetized randomly inhomogeneous plasma. The importance of the theory of transition radiation for astrophysics is illustrated by one example of its detailed application to a solar radio burst, including specially designed algorithms of the spectral forward fitting.} \\def\\gsim{\\ \\raise 3pt \\hbox{$>$} \\kern -8.5pt \\raise -2pt \\hbox{$\\sim$}\\ } \\def\\lsim{\\ \\raise 3pt \\hbox{$<$} \\kern -8.5pt \\raise -2pt \\hbox{$\\sim$}\\ } ", "introduction": "The phenomenon of transition radiation was discovered theoretically by two Nobel Prize winning (2003 and 1958 respectively) physicists \\cite{Gin_Fr}. Ginzburg and Frank (1946) considered a simplest case when a charged particle passed through a boundary between two dielectrically different media and so generated waves due to a variation of the dielectric constant at the boundary. Remarkably, no acceleration of the particle is necessary to produce the emission due to transition through the boundary. It is easy to understand that a similar effect of electromagnetic emission will take place if a medium is uniformly filled by turbulence that produces fluctuations of the dielectric constant throughout the whole volume rather than at an isolated boundary. Many astrophysical sources, especially those under strong energy release, are believed to be filled by turbulent, randomly inhomogeneous plasma and fast, nonthermal particles. In this situation, an efficient contribution of the transition radiation to the overall electromagnetic emission should be produced. Therefore, distinguishing this contribution from competing mechanisms is important. Below we describe the fundamentals of the transition radiation produced in a magnetized turbulent plasma, and demonstrate its high potential for astrophysical applications. ", "conclusions": "Nita et al. (2005) proved that the dm continuum component of this solar radio burst is produced by RTR and derived the level of the microturbulence in the plasma to be $\\left<\\Delta n^2\\right>/n^2 = 10^{-5}$. This finding is potentially very important for other cosmic objects. Indeed, the obtained microturbulence level is not particularly strong and much stronger turbulence is expected in many cases, especially, when there is a strong release of the energy at the source. Sometimes, such energy release gives rise to a relativistic expansion of the source, so the emission spectrum is Doppler-boosted and RTR produced at the local plasma frequency can be observed at the Earth even from relatively tenuous sources with low plasma frequency. In this study we present more evidence in favor of RTR generation at the dm continuum solar bursts and use this emission component to derive additional plasma parameters. In particular, we determine the mean plasma frequency and its dispersion at the source in the course of time. Interestingly, these two parameters do not change much during the time of the dm burst. We note that these parameters are obtained from the total power spectra recorded without spatial resolution. Fig. 2 demonstrates that the radio sources at various frequencies do not coincide exactly. Therefore, in cases where a sequence of spatially resolved spectra are available we would be able to study the structure of the flaring plasma density in much greater detail as well as the distribution of the microturbulence over the source. Generally speaking, the RTR contribution is also informative about the fast electrons producing it. In the example presented in Fig. 3 we show the emission decay constants, which can be associated with the fast electron life times. In our case we used exponential fragments of the light curves at the late decay phase of the emission, since no exponential phase was found in the early decay phase. We found the life time to be within 10-40 sec, which corresponds to the electrons of 300 keV or larger in the case of dense flare plasma available in this event. On the other hand, we can expect that most of the RTR emission (around the peak of the burst) is produced by the electrons with E=100-200 keV (Nita et al. 2005). This apparent contradiction can be easily resolved if we recall that the lower energy electrons have the life time of only a few seconds in the given dense plasma, so they die even before the light curves reach the exponential decay stage, and we observe the RTR contribution from preferentially higher energy electrons late in the event." }, "0710/0710.5574_arXiv.txt": { "abstract": "We study the effect of quasar feedback on distributions of baryons and properties of intracluster medium in galaxy groups using high-resolution numerical simulations. We use the entropy-conserving Gadget code that includes gas cooling and star formation, modified to include a physically-based model of quasar feedback. For a sample of ten galaxy group-sized dark matter halos with masses in the range of $1$ to $5\\times 10^{13} M_{\\odot}/h$, star formation is suppressed by more than 50\\% in the inner regions due to the additional pressure support by quasar feedback, while gas is driven from the inner region towards the outer region of the halos. As a result, the average gas density is 50\\% lower in the inner region and 10\\% higher in the outer region in the simulation, compared to a similar simulation with no quasar feedback. Gas pressure is lowered by about 40\\% in the inner region and higher in the outer region, while temperature and entropy are enhanced in the inner region by about 20-40\\%. The total group gas fraction in the two simulations generally differs by less than 10\\%. We also find a small change of the total thermal Sunyaev-Zeldovich distortion, leading to 10\\% changes in the microwave angular power spectrum at angular scales below two arcminutes. % ", "introduction": "Galaxy clusters and groups are the largest gravitationally bound objects in the universe, and they dominate the total baryon content of the universe. Their spatial distribution and mass function contain information about the formation and evolution of large-scale structure, which in turn constrain a variety of fundamental cosmological properties including normalization of the matter power spectrum, the cosmic baryon density, and dark matter properties. However, in order to use them as a cosmological probe, it is necessary to understand their astrophysical properties, and in particular their baryon physics. This issue is of particular current interest due to upcoming arcminute-resolution microwave sky surveys like ACT \\citep{kosowsky06,fowler07} and SPT \\citep{ruhl05}, which will image galaxy clusters via the Sunyaev-Zeldovich distortions to the cosmic microwave blackbody spectrum from the hot electrons in the cluster gas \\citep{sz80}. The majority of baryons in clusters and groups are in the form of hot intracluster gas rather than than individual galaxies. Properties of the Intracluster Medium (ICM) have been studied through a combination of X-ray and radio observations \\citep{nulsen05, heinz02, fabian2000}. Although the dark matter distribution in galaxy clusters follow a self-similar relation \\citep{pointecouteu05,vikhilin06}, the hot gas does not \\citep{sanderson03,popesso05}. Additional non-gravitational sources of heating are required to explain the observations. One interesting and plausible possibility is the energy radiated from quasars or Active Galactic Nuclei (AGN) and deposited into the ICM \\citep{kaiser91,valageassilk99,nath02, evan05, evan06}, which we study in this work. The best arena in which to study the impact of various feedback mechanisms is galaxy groups. Massive clusters with deeper gravitational potential wells are likely to have their global thermodynamic and morphological properties less affected by feedback. In comparison, galaxy groups have shallower potential wells while still having enough gas to display the effect of feedback on the ICM. Galaxy groups have recently been observed in X-rays at redshifts as large as $z=0.6$ \\citep{willis05}. In the optical band, \\cite{sdss07} have compiled group catalogs from the SDSS Data Release 5 catalog. Evidence for heating by a central AGN or radio source in galaxy groups and clusters has been the subject of several recent papers \\citep{croston05, jetha06,sanderson05}. These observations show excess entropy in cluster cores, which suggests that some heating process must act to offset cooling. In recent years, cosmological simulations including dark matter and gas have been able to follow the evolution of individual galaxy groups and clusters. A number of studies have investigated the cluster baryon fraction and its evolution in numerical simulations. Adiabatic simulations that do not include radiative cooling find cluster baryon fractions around $0.85$ of the universal baryon fraction \\cite{evrard90,metzler_evrard94,navarro_etal95,lubin_etal96,eke_etal98,frenk_etal99,mohr_etal99,bialek_etal01}. Preheating the gas reduces the fraction further \\citep{bialek_etal01,borgani_etal02,muanwong_etal02,kay_etal03}. When cooling, star formation and other feedback processes are included, the baryon fraction is higher than that obtained from adiabatic simulations \\citep{muanwong_etal02,kay_etal03,valdarnini03,ettori04,nagai07}. This leads to an ``overcooling'' problem and indicates an additional feedback mechanism. In the current study, we analyze the effect of quasar feedback on the baryon distribution and thermodynamics of hot gas in galaxy groups. We also study its implication for the Sunyaev-Zeldovich angular power spectrum, which receives a dominant contribution from high-redshift halos. Ref.~\\cite{ks02} showed that the thermal SZ angular power spectrum provides a strong constraint on the normalization of the matter power spectrum, $\\sigma_8$. Upcoming SZ surveys like ACT or SPT will have sufficient sensitivity to determine $\\sigma_8$ with an accuracy limited by uncertainty in the theoretical model. Also, the kinematic SZ effect is a measure of bulk motions in the universe and may be a competitive probe for studying cosmology \\citep{sehgal04, bk06, penn06, dedeo05, maturi07, roncarelli07}. But one of the major sources of uncertainty in modeling the kSZ effect is the gas fraction and its evolution. So understanding both the thermal and kinematic SZ signals requires detailed understanding of feedback mechanisms in galaxy clusters and groups. The mechanisms and effects of feedback are also a long-standing question in astrophysics, with particular bearing on the process of galaxy formation. To this end, we have analyzed a sample of ten galaxy groups at $z=1$ from numerical cosmological simulations of gas and dark matter which have been extended to include a self-consistent model for the evolution of massive black holes and their baryon feedback. At redshift $z>1$, the quasar mode of black hole accretion is expected to be the dominant feedback mechanism, compared to the radio-loud accretion mode which becomes important at lower redshifts \\citep{sijacki07}. The size of our simulations prevents studying feedback in galaxy clusters, but rather restricts us to less massive galaxy groups. But as already mentioned, galaxy groups with shallow potential wells provide the best place to study non-gravitational heating and its implications for the properties of hot gas. High-redshift galaxy groups are also a major contributor to the thermal SZ power spectrum, which peaks around $z\\approx 1$, when galaxy groups are more numerous than massive clusters \\cite{ks02}. Following this introduction, Section II describes our simulation and its implementation of quasar feedback. In Section III we study the effect of numerical resolution on our results; in Section IV we describe our results and compare them with a simulation that do not include quasar feedback. Finally, in Section V we summarize our results and discuss directions for future work, including motivations and prospects for studying more massive galaxy clusters and more realistic feedback modeling for quasars and AGN. ", "conclusions": " 1. Compared to the no-feedback case, star formation is suppressed by 30-40\\% in the inner regions of the halos because of the additional pressure support provided by quasar feedback. 2. Quasar feedback redistributes hot gas, driving it from the inner region towards the outer part of the halos. As a result, gas density is 20\\% less in the inner part and 10\\% to 15\\% greater in the outer region when compared to the simulation without feedback. However, the gas fraction in the two simulation differs by only 5\\% to 10\\%, and gas fractions tends to increase mildly with increasing halo mass. 3. The ratio of gas mass to stellar mass increases by a factor of 3.5 in the simulation including quasar feedback and a factor of 2.5 in the simulation without quasar feedback in the region $0.2 R_{200m}5000$. We find little dependence of the SZ enhancement with halo mass. The effects of quasar feedback on the intracluster medium will be most evident in the group-sized haloes considered here, with their relatively shallow gravitational potential wells. Observationally, the most interesting haloes are larger in mass by a factor of ten, galaxy clusters: these are the haloes which are most readily detected via their SZ, X-ray, or optical signals. The gas fractions do not show any particular trend with increasing halo mass, and star fractions increase very weakly with mass over the halo mass range studied here. So it is reasonable to expect that the results of this work will hold for cluster-sized halos as well. Nevertheless, given the substantial impact of quasar feedback on various properties of the intracluster medium which the current study suggests, it is imperative to study cluster-sized halos as well. This requires larger-volume simulations, as the number density of clusters decreases with cluster mass. To this end, we are currently running a simulation of box size $50\\,{\\rm Mpc}/h$; results will be reported elsewhere. This is the first attempt to study the impact of quasar feedback on the baryon fraction and thermodynamics of the intracluster medium in a cosmological hydrodynamic simulation. Both the gas and star fractions in our simulation are consistent with current observational limits \\citep{allen_etal04, ettori99}. Note that we have studied only quasar feedback at redshifts greater than unity. However, active galactic nuclei also inject energy into the ICM via a ``radio mode'' which is believed to be the dominant feedback mechanism at lower redshift \\citep{sijacki05, sijacki07}. Thus our results should be treated as a conservative estimate of the total impact of AGN feedback for galaxy groups at low redshifts. Gas pressure in cosmological halos, particularly those with masses ranging from galaxy groups to galaxy clusters, determines the important thermal Sunyaev-Zeldovich signal which will soon be measured with high precision. The gas fraction is important for connecting kinematic SZ signals of cluster gas momentum with theoretical predictions about cluster velocity or total momentum. This paper takes the first step towards quantifying the impact of quasars on these quantities, which turns out to be significant but not dominating. Much work remains to be done, both through larger simulations which contain many galaxy-cluster-sized haloes, and in enhancing the realism of the quasar feedback models. We hope the results here plus the exciting observational prospects in the near future will open the door to further advances in this area." }, "0710/0710.2481_arXiv.txt": { "abstract": "The stellar populations in the bulges of S0s, together with the galaxies' dynamics, masses and globular clusters, contain very interesting clues about their formation. I present here recent evidence suggesting that S0s are the descendants of fading spirals whose star formation ceased. ", "introduction": "Combining published data with high-quality VLT/FORS spectroscopy of sample of Fornax S0s (Bedregal et al.\\ 2006a) we have carried out a combined study of the Tully-Fisher relation and the stellar populations of these galaxies. Despite the relatively small sample and the considerable technical challenges involved in determining the true rotation velocity $V_{\\rm rot}$ from absorption line spectra of galaxies with significant non-rotational support (see Mathieu et al.\\ 2002), some very interesting results arise. S0s lie systematically below the spiral galaxy Tully-Fisher relation in both the optical and near-infrared (Figure~1). If S0s are the descendants of spiral galaxies, this offset can be naturally interpreted as arising from the luminosity evolution of spiral galaxies that have faded since ceasing star formation. Moreover, the amount of fading implied by the offset of individual S0s from the spiral relation seems to correlate with the luminosity-weighted age of their stellar population, particularly at their centres (Figure~2). This correlation suggests a scenario in which the star formation clock stopped when gas was stripped out from a spiral galaxy and it began to fade into an S0. The stronger correlation at small radii indicates a final last-gasp burst of star formation in this region. See Bedregal, Arag\\'on-Salamanca \\& Merrifield (2006b) for details. \\begin{figure} \\begin{center} \\includegraphics[height=3.3in,width=4.0in,angle=0]{Aragon-Salamanca_fig1.ps} \\end{center} \\caption{$B$-band Tully-Fisher relation (TFR) for S0 galaxies using different samples from the literature (open symbols) and our VLT Fornax data (filled circles). The solid and dashed lines show two independent determinations of the TFR relation for local spirals. On average (dotted line), S0s are $\\sim3$ times fainter than spirals at similar rotation velocities (Bedregal, Arag\\'on-Salamanca \\& Merrifield 2006b). } \\label{fig:fig1} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[height=1.65in,angle=0]{Aragon-Salamanca_fig2.eps} \\end{center} \\caption{ For our VLT Fornax data we plot the shift in magnitudes from the $B$-band spiral TFR versus the stellar population age at the galaxy centre (left panel), at $1\\,R_e$ (middle panel) and at $2\\,R_e$ (right panel). The lines show models for fading spirals. Note that the correlation is strongest for the central stellar populations of the galaxies, suggesting that the last episode of star formation took place there (Bedregal, Arag\\'on-Salamanca \\& Merrifield 2006b). } \\label{fig:fig2} \\end{figure} ", "conclusions": "The stellar populations, dynamics and globular clusters of S0s provide evidence consistent with these galaxies being the descendants of fading spirals whose star formation ceased. However, caution is needed since significant problems could still exist with this picture (see, e.g., Christlein \\& Zabludoff 2004; Boselli \\& Gavazzi 2006). Moreover, the number of galaxies studied here is still small, and it would be highly desirable to extend this kind of studies to much larger samples covering a broad range of galaxy masses and environments. \\begin{figure} \\begin{center} \\includegraphics[height=2.9in,angle=0]{Aragon-Salamanca_fig3.eps} \\end{center} \\caption{ Log$_{10}$ of the luminosity-weighted ages is Gyr vs.\\ the globular cluster specific frequency ($S_N$) of S0s. The line shows the evolution expected for a fading galaxy according to the stellar population models of Bruzual \\& Charlot (2003). The correlation between the fading of the galaxies (or increase in $S_N$) and the spectroscopically-determined age of their stellar populations is clearly consistent with the predictions of a simple fading model. Note that the $S_N$ value for NGC3115B is very unreliable and almost certainly severely overestimated due to contamination from the GC systems of neighbouring galaxies. See Barr et al.\\ (2007) for details. } \\label{fig:fig3} \\end{figure}" }, "0710/0710.0484_arXiv.txt": { "abstract": "We study the propagation of relativistic jets originating from AGNs within the Interstellar/Intergalactic Medium of their host galaxies, and use it to build a model for the suppression of stellar formation within the expanding cocoon. ", "introduction": "\\label{sec:jetcocoon} Relativistic jets from Supermassive Black Holes hosted within the bulges of spirals and nuclei of early-type galaxies inject significant amounts of energy into the medium within which they propagate, \\begin{figure} \\centering \\includegraphics[scale=0.6,angle=90]{antonucciodelogu_fig1.eps} \\caption{Expansion of the cocoon within the ISM. The jet has an input mechanical power $\\textrm{P}_{jet} = 10^{46} \\textrm{ergs}\\cdot\\textrm{cm}^{-3}\\cdot\\textrm{sec}^{-1}$, and the ISM density is: $\\, n_{e}^{ism} = 1 e^{-} cm^{-3}$.}\\label{fig:cocoon} \\end{figure} creating an extended, underdense and hot cocoon. We have performed a series of simulations of these jets and cocoons using an AMR code, FLASH 2.5: a typical output is shown in fig.~\\ref{fig:cocoon}.\\\\ We find that a slight modification of the exact models of \\cite{1991MNRAS.250..581F} and \\cite{1997MNRAS.286..215K}, approximates well the simulations. ", "conclusions": "" }, "0710/0710.2815_arXiv.txt": { "abstract": "We present evolutionary tracks of binary systems with high mass companion stars and stellar-through-intermediate mass BHs. Using Eggleton's stellar evolution code, we compute the luminosity produced by accretion from the donor during its entire evolution. We compute also the evolution of the optical spectrum of the binary system taking the disc contribution and irradiation effects into account. The calculations presented here can be used to constrain the properties of the donor stars in Ultraluminous X-ray Sources by comparing their position on the HR or color-magnitude diagrams with the evolutionary tracks of massive BH binaries. This approach may actually provide interesting clues also on the properties of the binary system itself, including the BH mass. We found that, on the basis of their position on the color-magnitude diagram, some of the candidate counterparts considered can be ruled out and more stringent constraints can be applied to the donor masses. ", "introduction": "Point like off nuclear X-ray sources with luminosities well in excess of the Eddington limit for a stellar mass black hole, have been discovered in a large number of nearby galaxies (e.g.~\\citealt{2002ApJS..143...25C}; \\citealt{sgtw04}; \\citealt{2005ApJS..157...59L}). These Ultraluminous X-ray Sources (ULXs) are too dim to be low luminosity AGNs and too bright to be normal X-ray binaries (XRBs) emitting below the Eddington limit. In this paper we define a ULX as a source with bolometric luminosity in excess of the Eddington limit for a stellar mass black hole of $20\\msun$ and less luminous than a few $10^{41}\\rm erg/s$. This definition implies that a Galactic XRB radiating isotropically at or below the Eddington limit can not be a ULX. In spiral or starburst galaxies ULXs turn out to be associated with star forming regions, emission nebulae and stellar clusters (\\citealt{zfrm02}, \\citealt{pm02}). These facts along with, in some case, the detection of stellar optical counterparts (\\citealt{2001MNRAS.325L...7R}; \\citealt{2002MNRAS.335L..67G}; \\citealt{2002ApJ...580L..31L}; \\citealt{2004ApJ...602..249L};\\citealt{kwz04};\\citealt{zamp1}; \\citealt{2005ApJ...629..233K}; \\citealt{mucetal05, mucetal07}; \\citealt{scpmw04}) strongly indicate an association between ULXs and young massive stars, although the nature of the accreting compact object remains unclear. A certain number of ULXs, however, appear as isolated X-ray sources with no obvious counterpart at any wavelength (\\citealt{sm04}) and without a clear association with star forming regions or emission nebulae. Some low-luminosity ULXs may possibly be present also in elliptical galaxies (\\citealt{jcbg03}), but their nature and the actual evidence for their existence is not well established (see e.g. \\citealt{iba04}, \\citealt{aglc04}, \\citealt{2005ApJ...622L..89G}). The nature of the ULXs in spiral galaxies is less controversial. Different models have been proposed to explain the large luminosities reached by these sources. One of the favored models consists of an intermediate mass black hole (IMBH) with a mass in the range $10^{2}-10^{3}M_{\\rm \\odot}$, accreting from an high mass donor star. The presence of an IMBH can account for most of the observational properties of ULXs in a rather straightforward way. For instance, the observed cool disc spectra of some ULX can be explained with the fact that the innermost stable circular orbit of an IMBH is larger than that of a stellar mass black hole (e.g. \\citealt{mfm04}). The detection of a $\\sim$50-160 mHz quasi periodic oscillations in the power density spectrum of M82 X-1 and NGC5408 X-1 (\\citealt{2003ApJ...586L..61S}, \\citealt{2004ApJ...614L.113F}, \\citealt{2006MNRAS.365.1123M}; \\citealt{stroh07}), the very high luminosity of some ULXs ($\\sim 10^{41} \\ergs$) along with their cool discs, and the energy content and morphology of the nebulae around some of them (\\citealt{pm02}) all suggest an IMBH interpretation. The main problem with this interpretation resides in the formation mechanism of such an extreme object. In fact, if IMBHs with masses in excess of $\\sim 100 M_\\odot$ exist, they will require a new formation root with respect to the stellar black holes in our Galaxy and to the supermassive black holes in Active Galactic Nuclei. Until now two scenarios have been proposed to form a black hole in the intermediate mass range: the runaway collision of massive stars in dense open clusters (\\citealt{2004Natur.428..724P}, \\citealt{2004ApJ...604..632G}) and the primordial collapse of a very high mass star with zero metallicity (\\citealt{2000ApJ...540...39A}, \\citealt{2001ApJ...551L..27M}). Both the mechanisms however suffer of a certain degree of uncertainty related to the incomplete knowledge of the behavior of very massive stars. Therefore we have no final evidence that an IMBH can really form. Furthermore, the interpretation of the soft components observed in some ULXs in terms of cool accretion discs is not univocal (e.g. \\citealt{2005ApJ...631L.109C}, \\citealt{2005ApJ...635..198D}, \\citealt{2005MNRAS.357.1363R}, \\citealt{fk06}, \\citealt{gonc06}, \\citealt{2006MNRAS.368..397S}). Other interpretations in terms of stellar (or quasi-stellar) mass black holes have been proposed. A mechanical (\\citealt{kdwfe01}) or a relativistic beaming (\\citealt{kfm02}) can reproduce the observed luminosities of ULXs up to a few $10^{40}\\ergs$ with a beaming factor around $\\sim 10.$ At most, as suggested by \\citealt{2005MNRAS.357..275K}, accretion from helium rich matter from a geometrically thick disc can generate luminosities up to $\\sim 5\\times10^{40}\\ergs$. However, luminosities in excess of $5\\times 10^{40}\\ergs$ ($\\sim 5$\\% of the ULX population) and the isotropy of the ionized nebulae around some ULXs can not be easily explained in terms of beaming models. On the other hand, photon bubbles disc instabilities (\\citealt{b02}, \\citealt{b06}) and emission from a slim disc (\\citealt{wmm01}, \\citealt{ezkmw03}) can produce genuine isotropic super-Eddington luminosities around 10 times the Eddington limit. As shown by \\cite{2005MNRAS.356..401R}, a normal binary with a stellar mass black hole and a donor star with initial mass $\\simgreat 10\\mdot$, can in principle explain a large sample of ULXs if a super Eddington luminosity around $10$ is allowed (see also \\citealt{2003MNRAS.341..385P}). However, the typical temperature of a slim disc in this regime ($\\sim$ 1--2 keV) is inconsistent with the observation of some cool disc sources which are, at the same time, the most luminous ULXs. Furthermore, luminosities in excess of a few $10^{40}\\ergs$ are not attainable with slim discs around stellar mass black holes. On the other hand, the photon bubble instability model makes no predictions on the observable accretion disc temperature and therefore cannot be compared directly with observations. Thanks to the high precision astrometry of the {\\it Chandra} X-ray observatory, we know that some ULXs have optical counterparts which match with the X-ray source position (\\citealt{2002ApJ...580L..31L}, \\citealt{2004ApJ...602..249L}, \\citealt{zamp1}, \\citealt{kwz04}, \\citealt{sm04}, \\citealt{2005ApJ...629..233K}, \\citealt{mucetal05, mucetal07}, \\citealt{Liu07}). Most of them are suspected to be main sequence (MS) high mass stars or supergiant stars of uncertain spectral type. The ambiguity arises because, until now, optical spectra of these stars either are not available, or they are too noisy and with peculiar spectral features (\\citealt{2004ApJ...602..249L}), or there is more than an optical counterpart in the X-ray error box (\\citealt{sm04}, \\citealt{mucetal05}). In this paper, we present the evolutionary tracks of binary systems with high mass companion stars and stellar-through-intermediate mass BHs and compare them with the properties of the donor stars in ULX binary systems. In \\S 2 we present the model adopted to evolve the binary system. In \\S 3, we summarize the properties of the four ULXs considered and their optical counterparts. Our results are presented in \\S 4 and compared with the properties of ULX counterparts in \\S 5. Conclusions follow in \\S 6. ", "conclusions": "In this paper we have shown that the effects of the binary evolution on the optical properties of a donor star in a ULX binary system can significantly alter its colors and evolution. We included also the contribution of the optical emission of the accretion disc and the X-ray irradiation of the donor and disc surfaces. The evolutionary track of a donor in an accreting binary on the CM diagram is very different with respect to that of a single isolated star. These important differences can not be overlooked when trying to identify the donor mass of a ULX. We calculated tracks for stellar and intermediate mass black holes, and demonstrate the brighter nature of the IMBH as an intrinsic phenomenon produced by the low mass ratio in the binary. The photometric data of four ULX counterparts have been compared with the evolutionary tracks of massive donors on the CM diagram. We find that the counterparts 2, 3 and 4 of NGC 4559 X-7 are consistent only with donors of $10$--$15\\msun$ undergoing a case B or C mass transfer episode. The only possibility for objects 5 and 8 is a very massive ($\\sim50\\msun$) MS donor or a H-shell burning donor with mass between 10-15 and 30$\\msun$. Finally, object 1 is compatible only with a very massive ($M\\simgreat 50\\msun$) donor in a H-shell burning phase. Our results are quite different with respect to what obtained in previous analysis. \\cite{scpmw04} compared theoretical evolutionary tracks of single stars between 9 and 25$\\msun$ with the position in the CM diagram of the counterparts of NGC4559 X-7. They find that the colors and luminosity of six out of seven stars are consistent with the tracks of main sequence, blue or red supergiant with masses in the range 10-15 $M_\\odot$ and ages $\\sim 20$ Myr. Their favored counterpart, object 1, was identified with a main sequence or blue supergiant of $\\sim 20 M_\\odot$ and an age of $\\sim 10$ Myr, although \\cite{scpmw04} recognize that X-ray irradiation may affect its colors and hence its classification. Irradiation effects were taken into account by \\cite{cpw07}, giving a donor mass of 5-20$M_{\\odot}$ for object $1$. In their work, however, \\cite{cpw07} considered irradiation on single stars and did not take into account the effects of the markedly different evolution of a donor star in a binary system. As far as NGC1313 X-2 is concerned, we definitely rule out the candidate counterpart C2 and identify C1 as a H-shell burning donor of 10$\\msun<$M$<$15$\\msun$ around a stellar mass black hole or a $\\simless 15 \\msun$ donor undergoing RLOF during MS (or the H-shell burning phase) in an IMBH binary system. On statistical grounds alone, it is then more likely that NGC1313 X-2 hosts a $\\sim 100 \\msun$ BH rather than a stellar mass BH as the H-shell burning phase is much shorter than the MS phase. Our result for object C2 leaves only C1 as the likely counterpart of NGC 1313 X-2, in agreement with the evidence coming from the refined X-ray astrometry of the field recently reported by \\cite{Liu07}. On the basis of single star isochrone fitting of the parent stellar population, \\cite{pak06} and \\cite{ramsey06} estimate a maximum MS mass of 8--9$\\msun$ for object C1. \\cite{Liu07} find consistency with either a $\\sim 8\\msun$ star of very low metallicity or an O spectral type, solar metellicity star of $\\sim 30\\msun$. Our estimated range of donor masses for object C1 appears more in agreement with the value (10-18$\\msun$) reported by \\cite{mucetal07}, probably because they also included the contribution of the disc and irradiation effects. In Holmberg II X-1 there are several possibilities, and we can only rule out stars with M$<$10$\\msun$ (\\citealt{cpw07} give a lower bound of 5$\\msun$), while for M81 X-9 we are left with the unique possibility that the donor is a $\\sim 10\\msun$ star in the H-shell burning phase. We conclude therefore that in all previous work where the binary evolution and/or irradiation effects ware not taken into account, the donor mass has been systematically overestimated. We find that mass transfer occurring at late stages is very unlikely, not only for the short timescale of the contact phase ($10^{3}$--$10^{5}$yrs) but also for the incompatibility of the optical colors with the majority of the observed counterparts. If mass transfer sets in during MS, two contact phases occur (segments A--B and C--D). The mass transfer timescales are $t_{A-B}\\sim 10^{6}-10^{7}\\rm\\,yrs$ and $t_{C-D}\\sim 10^{3}-10^{5}\\rm\\,yrs$ (depending on the donor mass). Therefore assuming a flat distribution of periods between 1 and $\\sim$10 days, as during MS the contact phase is reached for $P_{orb}\\simless 2$days, we expect a factor $\\sim \\frac{10}{2}\\times \\frac{t_{A-B}}{t_{C-D}}\\simeq 500$--$5000$ more main sequence systems than H-shell burning donors." }, "0710/0710.4328_arXiv.txt": { "abstract": "The initial conditions and relevant physics for the formation of the earliest galaxies are well specified in the concordance cosmology. Using ab initio cosmological Eulerian adaptive mesh refinement radiation hydrodynamical calculations, we discuss how very massive stars start the process of cosmological reionization. The models include non-equilibrium primordial gas chemistry and cooling processes and accurate radiation transport in the Case B approximation using adaptively ray traced photon packages, retaining the time derivative in the transport equation. Supernova feedback is modeled by thermal explosions triggered at parsec scales. All calculations resolve the local Jeans length by at least 16 grid cells at all times and as such cover a spatial dynamic range of $\\sim$10$^6$. These first sources of reionization are highly intermittent and anisotropic and first photoionize the small scales voids surrounding the halos they form in, rather than the dense filaments they are embedded in. As the merging objects form larger, dwarf sized galaxies, the escape fraction of UV radiation decreases and the \\ion{H}{2} regions only break out on some sides of the galaxies making them even more anisotropic. In three cases, SN blast waves induce star formation in overdense regions that were formed earlier from ionization front instabilities. These stars form tens of parsecs away from the center of their parent DM halo. Approximately 5 ionizing photons are needed per sustained ionization when star formation in 10$^6$ \\Ms~halos are dominant in the calculation. As the halos become larger than $\\sim$$10^7 \\Ms$, the ionizing photon escape fraction decreases, which in turn increases the number of photons per ionization to 15--50, in calculations with stellar feedback only. Radiative feedback decreases clumping factors by 25 per cent when compared to simulations without star formation and increases the average temperature of ionized gas to values between 3,000 and 10,000 K. ", "introduction": "It is clear that quasars are not responsible to keep the universe ionized at redshift 6. The very brightest galaxies at those redshifts alone also provide few photons. The dominant sources of reionization so far are observationally unknown despite remarkable advances in finding sources at high redshift \\citep[e.g.][]{Shapiro86, Bouwens04, Fan06, Thompson07, Eyles06} and hints for a large number of unresolved sources at very high redshifts \\citep{Spergel07, Kashlinsky07} which is still a topic of debate \\citep{Cooray07, Thompson07}. At the same time, ab initio numerical simulations of structure formation in the concordance model of structure formation have found that the first luminous objects in the universe are formed inside of cold dark matter (CDM) dominated halos of total masses $2 \\times 10^5 - 10^6 \\Ms$ \\citep{Haiman96, Tegmark97, Abel98}. Fully cosmological ab initio calculations of \\citet{Abel00, Abel02} and more recently \\citet{Yoshida06} clearly show that these objects will form isolated very massive stars. Such stars will be copious emitters of ultraviolet (UV) radiation and are as such prime suspect to get the process of cosmological reionization started. In fact, one dimensional calculations of \\citet{Whalen04} and \\citet{Kitayama04} have already argued that the earliest \\ion{H}{2} regions will evaporate the gas from the host halos and that in fact most of the UV radiation of such stars would escape into the intergalactic medium. Recently, \\citet{Yoshida07a} and \\citet{Abel07} demonstrated with full three-dimensional radiation hydrodynamical simulations that indeed the first \\ion{H}{2} regions break out of their host halos quickly and fully disrupt the gaseous component of the cosmological parent halo. All of this gas finds itself radially moving away from the star at $\\sim30\\kms$ at a distance of $\\sim100$ pc at the end of the stars life. At this time, the photo-ionized regions have now high electron fractions and little destructive Lyman-Werner band radiation fields creating ideal conditions for molecular hydrogen formation which may in fact stimulate further star formation above levels that would have occurred without the pre-ionization. Such conclusion have been obtained in calculations with approximations to multi dimensional radiative transfer or one dimensional numerical models \\citep{Ricotti02a, Nagakura05, OShea05, Yoshida06, Ahn07, Johnson07}. These early stars may also explode in supernovae and rapidly enrich the surrounding material with heavy elements, deposit kinetic energy and entropy to the gas out of which subsequent structure is to form. This illustrates some of the complex interplay of star formation, primordial gas chemistry, radiative and supernova feedback and readily explains why any reliable results will only be obtained using full ab initio three dimensional hydrodynamical simulations. In this paper, we present the most detailed such calculations yet carried out to date and discuss issues important to the understanding of the process of cosmological reionization. It is timely to develop direct numerical models of early structure formation and cosmological reionization as considerable efforts are underway to \\begin{enumerate} \\item Observationally find the earliest galaxies with the James Webb Space Telescope \\citep[JWST;][]{Gardner06} and the Atacama Large Millimeter Array \\citep[ALMA;][]{Wilson05}, \\item Further constrain the amount and spatial non-uniformity of the polarization of the cosmic microwave background radiation \\citep{Page07}, \\item Measure the surface of reionization with LOFAR \\citep{Rottgering06}, MWA \\citep{Bowman07}, GMRT \\citep{Swarup91} and the Square Kilometer Array \\citep[SKA;][]{Schilizzi04}, and \\item Find high redshift gamma ray bursts with SWIFT \\citep{Gehrels04} and their infrared follow up observations. \\end{enumerate} We begin by describing the cosmological simulations that include primordial star formation and accurate radiative transfer. In \\S\\ref{sec:SF}, we report the details of the star formation environments and host halos in our calculations. Then in \\S\\ref{sec:reion}, we describe the resulting start of cosmological reionization, and investigate the environments in which these primordial stars form and the evolution of the clumping factor. We compare our results to previous calculations and further describe the nature of the primordial star formation and feedback in \\S\\ref{sec:discussion}. Finally we summarize our results in the last section. \\begin{deluxetable*}{lccccccc} \\tablecolumns{8} \\tabletypesize{} \\tablewidth{\\textwidth} \\tablecaption{Simulation Parameters\\label{tab:sims}} \\tablehead{ \\colhead{Name} & \\colhead{$l$} & \\colhead{Cooling model} & \\colhead{SF} & \\colhead{SNe} & \\colhead{N$_{\\rm{part}}$} & \\colhead{N$_{\\rm{grid}}$} & \\colhead{N$_{\\rm{cell}}$} \\\\ \\colhead{} & \\colhead{[Mpc]} & \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{} } \\startdata SimA-Adb & 1.0 & Adiabatic & No & No & 2.22 $\\times$ 10$^7$ & 30230 & 9.31 $\\times$ 10$^7$ (453$^3$) \\\\ SimA-HHe & 1.0 & H, He & No & No & 2.22 $\\times$ 10$^7$ & 40601 & 1.20 $\\times$ 10$^8$ (494$^3$) \\\\ SimA-RT & 1.0 & H, He, \\hh & Yes & No & 2.22 $\\times$ 10$^7$ & 44664 & 1.19 $\\times$ 10$^8$ (493$^3$) \\\\ SimB-Adb & 1.5 & Adiabatic & No & No & 1.26 $\\times$ 10$^7$ & 23227 & 6.47 $\\times$ 10$^7$ (402$^3$) \\\\ SimB-HHe & 1.5 & H, He & No & No & 1.26 $\\times$ 10$^7$ & 21409 & 6.51 $\\times$ 10$^7$ (402$^3$) \\\\ SimB-RT & 1.5 & H, He, \\hh & Yes & No & 1.26 $\\times$ 10$^7$ & 24013 & 6.54 $\\times$ 10$^7$ (403$^3$) \\\\ SimB-SN & 1.5 & H, He, \\hh & Yes & Yes & 1.26 $\\times$ 10$^7$ & 24996 & 6.39 $\\times$ 10$^7$ (400$^3$) \\enddata \\tablecomments{Col. (1): Simulation name. Col. (2): Box size. Col. (3): Cooling model. Col. (4): Star formation. Col. (5): Supernova feedback. Col. (6): Number of dark matter particles. Col. (7): Number of AMR grids. Col. (8): Number of unique grid cells.} \\end{deluxetable*} ", "conclusions": "\\label{sec:discussion} We have studied the details of massive metal-free star formation and its role in the start of cosmological reionization. We have treated star formation and radiation in a self-consistent manner, allowing for an accurate investigation of the evolution of cosmic structure under the influence of early Pop III stars. Stellar radiation from these stars provides thermal, dynamical, and ionizing feedback to the host halos and IGM. Although Pop III stars are not thought to provide the majority of ionizing photons needed for cosmological reionization, they play a key role in the early universe because early galaxies that form in these relic \\ion{H}{2} regions are significantly affected by Pop III feedback. Hence it is important to consider primordial stellar feedback while studying early galaxy formation. In this section, we compare our results to previous numerical simulations and semi-analytic models of reionization and then discuss any potential caveats of our methods and possible future directions of this line of research. \\begin{figure}[t] \\begin{center} \\epsscale{1.15} \\plotone{f14_color} \\caption{\\label{fig:filtering} The Jeans mass $M_J$ and filtering mass $M_F$ that can form bound objects. The squares denote the total mass of star forming halos in all three simulations.} \\end{center} \\end{figure} \\subsection{Comparison to Previous Models} \\subsubsection{Filtering Mass} \\label{sec:filtering} One source of negative feedback is the suppression of gas accretion into potential wells when the IGM is preheated. The lower limit of the mass of a star forming halo is the Jeans filtering mass \\begin{equation} \\label{eqn:filtering} M_F^{2/3}(a) = \\frac{3}{a} \\int_0^a \\> da^\\prime M_J^{2/3}(a^\\prime) \\left[ 1 - \\left(\\frac{a^\\prime}{a}\\right) \\right], \\end{equation} where $a$ and $M_J$ are the scale factor and time dependent Jeans mass in the \\ion{H}{2} region \\citep{Gnedin98, Gnedin00b}. Additionally, the virial shocks are weakened if the accreting gas is preheated and will reduce the collisional ionization in halos with $\\tvir \\gsim 10^4$ K. To illustrate the effect of Jeans smoothing, we take the large \\ion{H}{2} region of SimB-SN because it has the largest ionized filling fraction, which is constantly being heated after $z = 21$. Temperatures in this region fluctuates between 1,000~K and 30,000~K, depending on the proximity of the currently living stars. In Figure \\ref{fig:filtering}, we show the resulting filtering mass of regions with an ionization fraction greater than $10^{-3}$ along with the total mass of star forming halos. \\citet{Gnedin00b} found the minimum mass of a star forming halo is better described by $M_F$ instead of $M_J$. Our simulations are in excellent agreement for halos that are experiencing star formation after reincorporation of their previously expelled gas. The filtering mass is the appropriate choice for a minimum mass in this case as the halo forms from preheated gas. However for halos that have already assembled before they become embedded in a relic \\ion{H}{2} region, the appropriate minimum mass $M_{\\rm{min}}$ is one that is regulated by the LW background \\citep{Machacek01, Wise05} and photo-evaporation \\citep[e.g.][]{Efstathiou92, Barkana99, Haiman01, Mesigner06}. This is evident in the multitude of star forming halos below $M_F$. With the exception of star formation induced by SN blast waves or I-fronts, this verifies the justification of using $M_{\\rm{min}}$ and $M_F$ for Pop III and galaxy formation, respectively, as a criterion for star forming halos in semi-analytic models. \\begin{figure*}[t] \\begin{center} \\epsscale{1.15} \\plotone{f15} \\caption{\\label{fig:aniso} Density (left) and temperature (right) slices of an anisotropic \\ion{H}{2} region in the most massive halo of SimB-RT. The star has lived for 2.5 Myr out of its 2.7 Myr lifetime. The field of view is 900 proper parsecs.} \\end{center} \\end{figure*} \\subsubsection{Star Formation Efficiency} \\label{sec:SFeff} Semi-analytic models rely on a star formation efficiency $f_\\star$, which is the fraction of collapsed gas that forms stars, to calculate quantities such as emissivities, chemical enrichment, and IGM temperatures. Low-mass halos that form a central star have $f_\\star \\sim 10^{-3}$ whose value originates from a single 100~\\Ms~star forming in a dark matter halo of mass 10$^6$~\\Ms~\\citep{Abel02, Bromm02, Yoshida06}. Pop II star forming halos are usually calibrated with star formation efficiencies from local dwarf and high-redshift starburst galaxies and are usually on the order of a few percent \\citep[e.g.][]{Taylor99, Gnedin00a}. This leads to the question: how efficient is star formation in these high-redshift halos while explicitly considering feedback? This is especially important when halos start to form multiple massive stars and when metallicities are not sufficient to induce Pop II star formation. The critical metallicity for a transition to Pop II is still unclear. Recently, \\citet{Jappsen07a} showed that metal line cooling is dynamically unimportant in diffuse gas until metallicities of $10^{-2} \\> Z_\\odot$. On the other hand, dust that is produced in SNe can generate efficient cooling down in dense gas with $10^{-6} \\> Z_\\odot$ \\citep{Schneider06}. If the progenitors of the more massive halos did not result in a pair-instability SN, massive star formation can continue until it becomes sufficiently enriched. Hence our simulations can probe the efficiency of this scenario of massive metal-free star formation. It has also been suggested that the cosmological conditions that lead to the collapse of a metal-poor molecular cloud ($Z/Z_\\odot \\approx 10^{-3.5}$) may be more important than some critical metallicity in determining the initial mass function of a given stellar system \\citep{Jappsen07b}. We calculate $f_\\star$ with the ratio of the sum of the stellar masses to the total gas mass of unique star-forming halos. For example at the final redshift of 15.9 in SimA-RT, the most massive halo and its progenitors had hosted 11 stars and the gas mass of this halo is $1.8 \\times 10^6 \\Ms$, which results in $f_\\star = 6.1 \\times 10^{-4}$ for this particular halo. Expanding this quantity to all star forming halos, $f_\\star/10^{-4} = 5.6, 6.7, 7.4$ for SimA-RT, SimB-RT, and SimB-SN, respectively. We note that our choice of $M_\\star = 170 \\Ms$ in SimB-SN increases $f_\\star$ by 70\\%. Our efficiencies are smaller than the isolated Pop III case because halos cannot form any stars once the first star expels the gas, and 40 -- 75 million years must pass until star can form again when the gas is reincorporated into the halo. By regarding the feedback created by Pop III stars and associated complexities during the assembly of these halos, the $f_\\star$ values of $\\sim$$6 \\times 10^{-4}$ that are explicitly determined from our radiation hydrodynamical simulations provide a more accurate estimate on the early star formation efficiencies. \\subsubsection{Intermittent \\& Anisotropic Sources} Our treatment of star formation and feedback produces intermittent star formation, especially in low-mass halos. If one does not account for this, star formation rates might be overestimated in this phase of star formation. Kinetic energy feedback is the main cause of this behavior. As discussed in sections \\ref{sec:haloMass} and \\ref{sec:kinetic}, shock waves created by D-type I-fronts and SN explosions expel most of the gas in halos with masses $\\lsim 10^7$ \\Ms. A period of quiescence follows these instances of star formation. Then stars are able to form after enough material has accreted back into the halo. Only when the halo becomes massive enough to retain most of the outflows and cool efficiently through \\lya~and \\hh~radiative processes, star formation becomes more regular with successive stars forming. The central gas structures in the host halo are usually anisotropic as it is acquiring material through accretion along filaments and mergers. At scales smaller than 10 pc, the most optically thick regions produce shadows where the gas radially behind the dense clump is not photo-ionized or photo-heated by the source. This produces cometary and so-called elephant trunk structures that are also seen in local star forming regions and have been discussed in detail since \\citet{Pottasch58}. At a larger distance, the surrounding cosmic structure is composed of intersecting or adjacent filaments and satellite halos that breaks spherical symmetry. The filaments and nearby halos are optically thick and remain cool and thus the density structures are largely unchanged. The entropy of dense regions are not increased by stellar radiation and will feel little negative feedback from an entropy floor that only exists in the ionized IGM \\citep[cf.][]{Oh03}. Ray-tracing allows for accurate tracking of I-fronts in this inhomogeneous medium. Radiation propagates through the least optically thick path and generates champagne flows that have been studied extensively in the context of present day star formation \\citep[e.g.][]{Franco90, Churchwell02, Shu02, Arthur06}. In the context of massive primordial stars, these champagne flows spread into the voids and are impeded by the inflowing filaments. The resulting \\ion{H}{2} regions have ``butterfly'' morphologies \\citep{Abel99, Abel07, Alvarez06a, Mellema06, Yoshida07a}. We also point out that sources embedded in relic \\ion{H}{2} largely maintain or increase the ionization fraction. Here the already low optical depth of the recently ionized medium (within a recombination time) allows the radiation to travel to greater distances than a halo embedded in a completely neutral IGM. The \\ion{H}{2} regions become increasingly anisotropic in higher mass halos. We show an example of the morphology of a \\ion{H}{2} region near the end of the star's lifetime in a dark matter halo with mass $1.4 \\times 10^7 \\Ms$ in Figure \\ref{fig:aniso}. \\subsection{Potential Caveats and Future Directions} Although we have simulated the first generations of stars with radiation hydrodynamic simulations, our methods have neglected some potentially important processes and made an assumption about the Pop III stellar masses. One clear shortcoming of our simulations is the small volume and limited statistics of the objects studied here. However, it was our intention to focus on the effects of Pop III star formation on cosmological reionization and on the formation of an early dwarf galaxy instead of global statistics. The star formation only simulations (SimA-RT and SimB-RT) converge to the similar averaged quantities, e.g. ionized fraction, temperatures, star formation rates, at the final redshift. The evolution of these quantities differ because of the limited number of stars that form in the simulations, which then causes the evolution to depend on individual star formation times. This variance should be expected in the small volumes that we simulate and should not diminish the significance of our results. We have verified even in a 2.5-$\\sigma$ peak that Pop III stars cannot fully reionize the universe, which verified previous conclusions that low-luminosity galaxies provide the majority of ionizing photons. Furthermore, it is beneficial to study Pop III stellar feedback because it regulates the nature of star formation in these galaxies that form from pre-heated material. Further radiation hydrodynamics simulations of primordial star and galaxy formation with larger volumes while still resolving the first star forming halos of mass $\\sim$$3 \\times 10^5 \\Ms$ will improve the statistics of early star formation, especially in more typical overdensities, i.e. 1-$\\sigma$ peaks, some of which could survive to become dwarf spheroidal galaxies at $z = 0$. In this work, we treated the LW radiation field as optically thin, but in reality, \\hh~produces a non-zero optical depth above column densities of $10^{14} \\> \\mathrm{cm}^{-2}$ \\citep{Draine96}. Conversely, Doppler shifts of the LW lines arising from large velocity anisotropies and gradient may render \\hh~self-shielding unimportant up to column densities of $10^{20} - 10^{21} \\> \\mathrm{cm}^{-2}$ \\citep{Glover01}. If self-shielding is important, it will lead to increased star formation in low-mass halos even when a nearby source is shining. Moreover, \\hh~production can also be catalyzed ahead of I-fronts \\citep{Ricotti01, Ahn07}. In these halos, LW radiation will be absorbed before it can dissociate the central \\hh~core. On the same topic, we neglect any type of soft UV or LW background that is created by sources that are cosmologically nearby ($\\Delta z / z \\sim 0.1$). A soft UV background either creates positive or negative feedback, depending on its strength \\citep{Mesigner06}, and a LW background increases the minimum halo mass of a star-forming halo \\citep{Machacek01, Yoshida03, OShea07, Wise07b}. However in our calculations, the lack of self-shielding, which suppresses star formation in low-mass halos, and the neglect of a LW background, which allows star formation in these halos, may partially cancel each other. Hence one may expect no significant deviations in the SFRs and reionization history if one treats these processes explicitly. To address the incident radiation and the resulting UV background from more rare density fluctuations outside of our simulation volume, it will be useful to bridge the gap between the start of reionization on Mpc scales to larger scale (10 -- 100 Mpc) simulations of reionization, such as the work of \\citet{Sokasian03}, \\citet{Iliev06}, \\citet{Zahn07}, and \\citet{Kohler07}. Radiation characteristics from a volume that has similar overdensities as our Mpc-scale simulations can be sampled from such larger volumes to create a radiation background that inflicts the structures in our Mpc scale simulations. Inversely, perhaps the small-scale evolution of the clumping factor, filtering mass, and average temperature and ionization states can be used to create an accurate subgrid model in large volume reionization simulations. Another potential caveat is the continued use of primordial gas chemistry in metal enriched regions in the SN runs. Our simulations with SNe give excellent initial conditions to self-consistently treating the transition to low-mass star formation. In future work, we plan to introduce metal-line and dust cooling models \\citep[e.g. from][] {Glover07, Smith07} to study this transition. The one main assumption about Pop III stars in our calculations is the fixed, user-defined stellar mass. The initial mass function (IMF) of these stars is largely unknown, therefore we did not want to introduce an uncertainty by choosing a fiducial IMF. It is possible to calculate a rough estimate of the stellar mass by comparing the accretion rates and Kelvin-Helmholtz time of the contracting molecular cloud \\citep{Abel02, OShea05}. Protostellar models of primordial stars have also shown that the zero-age main sequence (ZAMS) is reached at 100 \\Ms~for typical accretion histories after the star halts its adiabatic contraction \\citep{Omukai03, Yoshida06}. Furthermore, we have neglected HD cooling, which may become important in halos embedded in relic \\ion{H}{2} regions and result in lower mass ($\\sim$$30 \\Ms$) metal-free stars \\citep{OShea05, Greif06, Yoshida07b}. Based on accretion histories of star forming halos, one can estimate the ZAMS stellar mass for each halo and create a more self-consistent and ab initio treatment of Pop III star formation and feedback. We conducted three radiation hydrodynamical, adaptive mesh refinement simulations that supplement our previous cosmological simulations that focused on the hydrodynamics and cooling during early galaxy formation. These new simulations concentrated on the formation and feedback of massive, metal-free stars. We used adaptive ray tracing to accurately track the resulting \\ion{H}{2} regions and followed the evolution of the photo-ionized and photo-heated IGM. We also explored on the details of early star formation in these simulations. Theories of early galaxy formation and reionization and large scale reionization simulations can benefit from the useful quantities and characteristics of the high redshift universe, such as SFR and IGM temperatures and ionization states, calculated in our simulations. The key results from this work are listed below. \\medskip 1. SFRs increase from $5 \\times 10^{-4}$ at redshift 30 to $6 \\times 10^{-3}$ \\sfr~at redshift 20 in our simulations. Afterwards the SFR begins to have a bursting nature in halos more massive than $10^7 \\Ms$ and fluctuates around $10^{-2}$ \\sfr. These rates are larger than the ones calculated in \\citet{Hernquist03} because our simulation volume samples a highly biased region that contains a 2.5-$\\sigma$ density fluctuation. The associated emissivity from these stars increase from 1 to $\\sim$100 ionizing photons per baryon per Hubble time between redshifts 15 and 30. 2. In order to provide a comparison to semi-analytic models, we calculate the star formation efficiency to be $\\sim$$6 \\times 10^{-4}$ averaged over all redshifts and the simulation volume. For Pop III star formation, this is a factor of two lower than stars that are not affected by feedback \\citep{Abel02, Bromm02, Yoshida06, OShea07}. 3. Shock waves created by D-type I-fronts expel most of the gas in the host halos below $\\sim$$5 \\times 10^6 \\Ms$. Above this mass, significant outflows that are still bound to the halo are generated. This feedback creates a dynamical picture of early structure formation, where star formation is suppressed in halos because of this baryon depletion, which is more effective than UV heating or the radiative dissociation of \\hh. 4. We see three instances of induced star formation in halos with masses $\\sim 3 \\times 10^6 \\Ms$. Here a star forms as a SN blast wave overtakes an overdensity created by an ionization front instability. \\hh~formation is catalyzed by additional free electrons in the relic \\ion{H}{2} region and in the SN blast wave \\citep{Ferrara98}. 5. As star formation occurs regularly in the simulation after redshift 25, four (six) ionizing photons are needed per sustained hydrogen ionization. As the most massive halo becomes larger than $\\sim$$10^7 \\Ms$ in the simulations without SNe, \\ion{H}{2} regions become trapped and ionizing radiation only escapes into the IGM in small solid angles. Hence the number of photons per effective ionization increases to 15 (50). In SimB-SN, stellar radiation from induced star formation have an escape fraction of nearly unity, which occur four times in the calculation. This allows the IGM to remain ionized at a volume fraction 3 times higher than without SNe. Similarly, the ionizing photon to ionization ratio also stays elevated at 10:1 instead of decreasing in the calculations with star formation only. 6. Our simulations that include star formation and \\hh~formation capture the entire evolution of the clumping factor that is used in semi-analytic models to calculate the effective enhancement of recombinations in the IGM. We showed that clumping factors in the ionized medium fluctuate around the 75\\% of the values found in adiabatic simulations. They evolve from unity at high redshifts and steadily increase to $\\sim$4 and 3.5 with and without SNe at $z = 17$, respectively. Photo-evaporation from stellar feedback causes the decrease of the clumping factor. 7. We calculated the Jeans filtering mass with the volume-averaged temperature only in fully and partially ionized regions, which yields a better estimate than the temperature averaged over both ionized and neutral regions. The filtering mass depends on the thermal history of the IGM, which mainly cools through Compton cooling. It increases by two orders of magnitude to $\\sim$$3 \\times 10^7 \\Ms$ at $z \\sim 15$. It describes the minimum mass a halo requires to collapse after hosting a Pop III star. For halos forming their first star, the minimum halo mass is regulated by the LW background \\citep{Machacek01} and photo-evaporation \\citep[e.g.][]{Haiman01}. \\medskip Pop III stellar feedback plays a key role in early star formation and the beginning of cosmological reionization. The shallow potential wells of their host halos only amplify their radiative feedback. Our understanding of the formation of the oldest galaxies and the characteristics of isolated dwarf galaxies may benefit from including the earliest stars and their feedback in galaxy formation models. Although these massive stars only partially reionized the universe, their feedback on the IGM and galaxies is crucial to include since it affects the characteristics of low-mass galaxies that are thought to be primarily responsible for cosmological reionization. Harnessing observational clues about reionization, observations of local dwarf spheroidal galaxies, and numerical simulations that accurately handle star formation and feedback may provide great insight on the formation of the first galaxies, their properties, and how they completed cosmological reionization." }, "0710/0710.4602_arXiv.txt": { "abstract": "We use semi-analytic techniques to evaluate the burst sensitivity of designs for the {\\it EXIST} hard X-ray survey mission. Applying these techniques to the mission design proposed for the Beyond Einstein program, we find that with its very large field-of-view and faint gamma-ray burst detection threshold, {\\it EXIST} will detect and localize approximately two bursts per day, a large fraction of which may be at high redshift. We estimate that {\\it EXIST}'s maximum sensitivity will be $\\sim 4$~times greater than that of {\\it Swift}'s Burst Alert Telescope. Bursts will be localized to better than 40~arcsec at threshold, with a burst position as good as a few arcsec for strong bursts. {\\it EXIST}'s combination of three different detector systems will provide spectra from 3~keV to more than 10~MeV. Thus, EXIST will enable a major leap in the understanding of bursts, their evolution, environment, and utility as cosmological probes. ", "introduction": "In its quest to find black holes throughout the universe, the {\\it Energetic X-ray Imaging Survey Telescope (EXIST)} will detect, localize and study a large number of gamma-ray bursts, events thought to result from the birth of stellar-mass black holes. We present the methods used to calculate {\\it EXIST}'s capabilities as a gamma-ray burst detector; we use the {\\it EXIST} design evaluated by the National Research Council's `Committee on NASA's Beyond Einstein Program: An Architecture for Implementation' (2007; see also Grindlay 2007). The combination of large detector area, broad energy coverage, and wide field-of-view (FOV) will result in the detection of a substantial number of bursts with a flux distribution extending to fainter fluxes than that of previous missions. Thus {\\it EXIST} should detect high redshift bursts, perhaps even bursts resulting from the death of Pop~III stars. {\\it EXIST}'s imaging detectors will localize the bursts, while the combination of detectors, both imaging and non-imaging, will result in well-determined spectra from 3~keV to well over 10~MeV. In this paper we first describe the {\\it EXIST} mission design (\\S 2), emphasizing aspects relevant to burst detection. Then we present the sensitivity methodology (\\S 3), which we apply to the individual coded mask sub-telescopes (\\S 4). {\\it EXIST} will consist of arrays of these detectors with overlapping FOVs, and the overall mission sensitivity results from adding the sensitivity of the individual sub-telescopes (\\S 5). Imaging using counts accumulated over different timescales increases the sensitivity (\\S 6). Finally, we combine these different calculations to evaluate {\\it EXIST}'s overall capabilities to study bursts (\\S 7). ", "conclusions": "From the preceding analysis, we can draw several conclusions on {\\it EXIST}'s impact on the study of gamma-ray bursts. First we estimate the {\\it EXIST} burst detection rate. The BATSE observations provide the cumulative burst rate as a function of the peak flux value $\\psi_B$ averaged over $\\Delta t=1$~s in the $\\Delta E=$50--300~keV band (Band 2002): \\begin{equation} N_B \\sim 550 \\left[ {{\\psi_B}\\over \\hbox{0.3 ph cm$^{-2}$ s$^{-1}$}} \\right]^{-0.8} \\hbox{ bursts yr$^{-1}$ sky$^{-1}$ } \\quad . \\end{equation} The HET threshold sensitivity for a single sub-telescope on-axis is $\\psi_B\\sim 0.12$~ph~cm$^{-2}$~s$^{-1}$ for $E_p > 100$~keV. Using the BATSE rate in eq.~10 and integrating over the solid angle distribution in Figure~6 gives a burst detection rate for the HET of $\\sim 400$ bursts per year. Note that this rate is over the BATSE-specific values of $\\Delta E$ and $\\Delta t$, and {\\it EXIST} will use at least two different values of $\\Delta E$ (see \\S 4.1) and a variety of $\\Delta t$ values (see \\S 6). Consequently this rate should be increased by approximately 50\\% to account for the soft, faint, long duration bursts to which BATSE was less sensitive than {\\it EXIST}'s HET will be; we therefore expect the HET array to detect $\\sim 600$~bursts per year. The value of $\\psi_B$ for an LET varies more with the burst spectral parameters than for an HET, and therefore estimates of the LET burst detection rate based on the BATSE rate are much more uncertain. For a single LET $\\psi_B\\sim 0.3$~ph~cm$^{-2}$~s$^{-1}$ on axis at $E_p=100$~keV, which gives a burst detection rate of $\\sim$180 bursts per year using eq.~10 and the LET distribution in Figure~6. This rate should be increased by a factor of 2 to account for the different energy band $\\Delta E$ and accumulation times $\\Delta t$. We use a larger adjustment factor for the LETs than for the HETs because the LETs' energy band will overlap less with BATSE's than the HETs'. We therefore expect the LET array to detect $\\sim350$~bursts per year. Next we simulate the spectra that the {\\it EXIST} suite of detectors will observe. Figure~9 shows a count spectrum (counts s$^{-1}$ keV$^{-1}$) for a moderately strong burst as it might be observed by the LETs (lefthand set of curves), HETs (middle set) and the CsI active shields for the HETs (righthand set; based on Garson et al. 2006a). The solid curves show the signal count rate, while the dashed curves provide the estimated background. Thus {\\it EXIST} will facilitate spectral-temporal studies. Particularly important to physical burst emission models is determining $E_p$, which is typically of order 250~keV (Kaneko et al. 2006). In addition, correlations of $E_p$ with other burst properties, such as the `isotropic' energy (the Amati relation---Amati 2006) or total energy (the Ghirlanda relation---Ghirlanda, Ghisellini \\& Lazzati 2004), have been proposed. `Pseudo-redshifts' calculated from the observables related to the burst-frame parameters in these relations can be used in burst studies when spectroscopic redshifts are not available, and can guide ground observers in allocating telescope time to observing potential high redshift bursts. The recently proposed Firmani relation (Firmani et al. 2006) correlates $E_p$, the peak luminosity, and a measure of the burst duration, all of which are related to observables in the gamma ray band. Thus pseudo-redshifts will be estimated using the Firmani relation based on {\\it EXIST} data alone, independent of observations by other facilities. With well determined broadband spectra down to 3~keV, {\\it EXIST} will be capable of determining whether the Band function (Band et al. 1993) suffices to describe burst spectra. For example, Preece et al. (1996) found evidence in the BATSE data for the presence of additional emission below 10~keV. By scaling from the EXIST survey's source localization (Grindlay 2007), we find that bursts should be localized at threshold by the HETs and LETs to better than 40~arcsec and 8~arcsec, respectively; this localization should scale as $\\sim(\\sigma-1)^{-1}$. Because the HETs are more sensitive to the LETs, the HET localization is relevant to the faintest bursts EXIST will detect. {\\it EXIST}'s burst capabilities calculated above will constitute a major leap beyond current detectors, and should increase the number of high redshift bursts detected. On average, high redshift bursts should be fainter, softer and longer than low redshift bursts (although the broad burst luminosity function and great variety in burst lightcurves and spectra obscure this trend). Figure~10 compares the detector sensitivities of the HET (solid curve) and LET (dashed curve) arrays to the BAT on {\\it Swift} (dot-dashed curve) and BATSE's Large Area Detector (LAD---dot-dot-dashed curve). As discussed above, the sensitivity is the threshold peak flux $F_T$ integrated over the 1--1000~keV band as a function of the spectrum's $E_p$; $\\alpha=-1$ and $\\beta=-2$ are assumed. In addition, the figure shows families of identical bursts at different redshifts (the curves with the points marked by `+'). Each family is defined by the value of $E_p$ in the burst frame; here again $\\alpha=-1$ and $\\beta=-2$ are assumed. In each family the burst would be observed to have $F_T$=7.5 ph cm$^{-2}$ s$^{-1}$ if it were at $z=1$. The points marked by `+' are spaced every $\\Delta z=1/2$; thus the uppermost points are at $z=1$ and the lowermost points are at $z=10$. The pulses in burst lightcurves become narrower (shorter) at higher energy, an effect that is generally proportional to $E^{0.4}$ (Fenimore et al. 1995). Since the observed lightcurve originated in a higher energy band, pulses should become narrower with redshift, reducing the peak flux when integrated over a fixed accumulation time; the plotted families include this effect. Finally, in \\S 6 we showed that forming images on long timescales increases the sensitivity to long duration bursts, as might result from cosmological time dilation." }, "0710/0710.1840_arXiv.txt": { "abstract": "{Detailed studies of Be stars in environments with different metallicities like the Magellanic Clouds or the Galactic bulge are necessary to understand the formation and evolution mechanisms of the circumstellar disks. However, a detailed study of Be stars in the direction of the bulge of our own galaxy has not been performed until now.} {The aim of this work is to report the first systematic search for Be star candidates in the direction of the Galactic Bulge. We present the full catalogue, give a brief description of the stellar variability seen, and show some light curve examples.} {We searched for stars matching specific criteria of magnitude, color and variability in the $I$ band. Our search was conducted on the 48 OGLE II fields of the Galactic Bulge. } {This search has resulted in 29053 Be star candidates, 198 of them showing periodic light variations. Nearly 1500 stars in this final sample are almost certainly Be stars, providing an ideal sample for spectroscopic multiobject follow-up studies. } {} ", "introduction": "Be stars are non-supergiant fast rotator B stars whose spectra have, or had at some time, one or more Balmer lines in emission (Collins 1987). This emission is originated from a flattened circumstellar disk and can come and go episodically on time scales of days to decades. The responsible mechanisms for the production and dynamics of the circumstellar gas are still not constrained. Possible mechanisms include non-radial pulsations, wind-compressed disk model, magnetic activity and binarity (Porter $\\&$ Rivinius 2003, and references therein).\\\\ Be stars are variable in brightness on three time scales that are often superimposed. Many of them (especially early type Be stars) show short-term photometric variability on time-scales of 0.2 to 2 days and amplitudes up 0.1 magnitudes, caused by non-radial pulsation or rotation (Percy et al. 2002, 2004). Some have mid-term variations on times scales form weeks to months, probably due to density waves within the disk (Sterken et al. 1996). Their amplitudes go to up to 0.2 magnitudes. They show also long-term variations from years to decades, with amplitudes up 0.8 magnitudes (Mennickent, Vogt, \\& Sterken 1994; Pavlovski et al. 1997; Hubert $\\&$ Floquet 1998; Percy $\\&$ Bakos 2001). Stagg (1987) found that this type of variability occurres in at least half of the Be stars. A few of Be stars are close binaries and other present ejection process due to magnetic activity resulting in outbursts (Hubert et al. 1997).\\\\ Many Galactic Be stars have been surveyed for photometric variability in order to detect and confirm short-term or long-term variations and to find correlations between them and get clues on the physical processes in Be stars. For instance, Hubert $\\&$ Floquet (1998) investigated the short-period variability of Be stars using analysis based on the Fourier and CLEAN algorithms on the Hipparcos photometry. Percy et al. (2002, 2004) analyzed a large sample of stars with Hipparcos photometry using a form of autocorrelation function. Typical problems in these studies are the gaps in the time distribution of the measurements and the limitations of the used algorithms.\\\\ In recent years, many Be-star like variables have been discovered in the Magellanic Clouds, showing a big variety of light curves, some of them reminding those of the Galactic Be stars and others never observed in that type of stars. Keller et al. (2002) concluded that most of these blue variables should be Be stars. Searches for Be stars in the Magellanic Clouds were performed by Mennickent et al. (2002), Keller et al. (1999, 2002) and Sabogal et al. (2005) on the basis of selection criteria applied to different photometric databases (OGLE-II and MACHO), and took into account amplitude of variability and ranges of color-magnitudes in the selection process. De Wit et al. (2006) investigated a subsample of the blue variables found by Mennickent et al. (2002) in the Small Magellanic Cloud and found that the photometric variability of these Be stars is due to variations in the amount of Brehmstrahlung due to the evolution of the circumstellar gas from a disk-shaped envelope towards a ring-like structure. \\\\ The study of Be stars is relevant to make contributions to several important branches of stellar physics. In particular, detailed studies of Be stars in environments with different metallicities like the Magellanic Clouds or the Galactic bulge is crucial to understand the formation and evolution mechanisms of the circumstellar disks. However, a detailed study of Be stars in the direction of the bulge of our own galaxy has not been performed until now.\\\\ A very large number of stars was observed in the region of the Galactic Bulge during the course of the second phase of the Optical Gravitational Lensing Experiment (OGLE II) (Udalski, Kubiak \\& Szymanski 1997). We have performed a search for Be star candidates into this database. Here we present the results of this search. ", "conclusions": "In this paper we have provided a catalogue of 29053 Be star candidates, 198 of them periodic, in the direction of the Galactic bulge that were selected basically on photometric criteria. Most of these Be star candidates are probably members of the Galactic disk and trace the gradient of metallicity towards the Galactic centre. They are ideal targets for future observing programs based on multiobject spectroscopy, narrow band photometry or H$\\alpha$ imaging surveys. These programs could eventually establish their true nature and broke the residual degeneracy with variable red giants in the red part of the (V-I) color distribution." }, "0710/0710.4524_arXiv.txt": { "abstract": "It is well know that the coronagraphic observations of halo CMEs are subject to projection effects. Viewing in the plane of the sky does not allow us to determine the crucial parameters defining geoeffectivness of CMEs, such as the velocity, width or source location. We assume that halo CMEs at the beginning phase of propagation have constant velocities, are symmetric and propagate with constant angular widths. Using these approximations and determining projected velocities and difference between times when CME appears on the opposite sides of the occulting disk we are able to get the necessary parameters. We present consideration for the whole halo CMEs from SOHO/LASCO catalog until the end of 2000. We show that the halo CMEs are in average much more faster and wider than the all CMEs from the SOHO/LASCO catalog. ", "introduction": "Space Weather is significantly controlled by coronal mass ejections (CMEs) which can affect the Earth in a different way. CMEs originating close to the central meridian, directed toward the Earth, excite the biggest scientific concern. In coronagraphic observations they appear as enhancement surrounding the entire occulting disk and they were called `halo CME'. Since the first identification by Howard et al. (1982) plenty of them were detected and now they are routinely recorded by the high sensitive SOHO/LASCO coronagraphs. In spite of large advantage over previous instruments, the SOHO/LASCO observations are still affected by a projection effect (Gopalswamy et al. 2000b). Viewing in the plane of the sky does not allow us to determine the crucial parameters defining geoeffectivness of CMEs, such as the velocity, width or source location. Prediction of the arrival of CME in the vicinity of Earth is critically important in space weather investigations. Basing on interplanetary shocks detected by Wind and the corresponding CMEs detected by SOHO, Gopalswamy et al. (2000a) developed and next (Gopalswamy, 2001) improved an empirical model to predict the arrival of CMEs at 1AU. The critical element affecting this model is the initial CME speed. The better prediction could be achieved if real initial velocities are used instead projected velocities determined from LASCO observations. Similarly, attempts made to estimate the projection effect based on the location of the solar source employ ad hoc assumptions on parameter such as the width of CMEs (Sheeley et al. 1999, Leblanc and Dulk, 2000). In the present paper we try to determine these crucial parameters defining geoeffectivness of CMEs, such as the velocity, width or source location. We assume that halo CMEs at the beginning phase of propagation have a constant velocities, are symmetric and propagate with constant angular widths. Using these approximations and determining projected velocities and difference between times when CME appears on the opposite sides of the occulting disk we are able to get necessary parameters. We present results for the whole halo CMEs from SOHO/LASCO catalog until the and of 2000. ", "conclusions": "In this paper we present possibility to estimate the crucial parameters determining geoeffectiveness of the halo CMEs. The clue of this method is based on the difference between times when the halo CME appears at the opposite sides (first and finally appearance) of the occulting disk. We considered the whole events form SOHO/LASCO catalog until the end of 2000. We were able to determine the real velocity, width and source location for 73th CMEs from our sample. Unfortunately, 58 events were symmetric or too faint to do necessary considerations. Results are listed in the three successive tables. This list could be use for further statistical examination or to prediction of the arrival of CME in the vicinity of Earth. Presented results suggest that the halo CMEs represent a specific class of CMEs which are very wide and fast. Using our results we have to remember that the simple model has several shortcomings: (i) CMEs may be accelerating, moving with constant speed or decelerating at the beginning phase of propagation. This means the constant velocity we assumed may not hold. (ii) CMEs may expand in addition to radial motion. Then the measured sky-plane speed is a sum of the expansion speed and the projected radial speed. This also would imply that the CMEs may not be a rigid cone as we assumed (Gopalswamy et al. 2001) (iii) The cone symmetry also may not hold. CME originating from loop structure could be elongated. All these limits can be overcome by stereoscopic observations only. Unfortunately, at the present time they are not available yet. It is necessary to develop the model to get the better fit to observations. The first step to improve our model could be achieved by consideration of acceleration and expansion of CMEs. \\\\ \\\\ {\\small \\bf Acknowledgments}\\\\ {\\scriptsize This paper was done during work of Grzegorz Michalek at Center for Solar and Space Weather, Catholic University of America in Washington.\\\\ In this paper we used data from SOHO/LASCO CME catalog. This CME catalog is generated and maintained by the Center for Solar Physics and Space Weather, The Catholic University of America in cooperation with the Naval Research Laboratory and NASA. SOHO is a project of international cooperation between ESA and NASA.\\\\ Work done by Grzegorz Michalek was partly supported by {\\it Komitet Bada\\'{n} Naukowych} through the grant PB 258/P03/99/17.} \\vspace{10mm}" }, "0710/0710.4238_arXiv.txt": { "abstract": "{Since most high- and intermediate-mass protostars are at great distance and form in clusters, high linear resolution observations are needed to investigate their physical properties.} {To study the gas in the innermost region around the protostars in the proto-cluster IRAS\\,05358+3543, we observed the source in several transitions of methanol and other molecular species with the Plateau de Bure Interferometer and the Submillimeter Array, reaching a linear resolution of 1100~AU.} {We determine the kinetic temperature of the gas around the protostars through an LVG and LTE analysis of their molecular emission; the column densities of CH$_3$OH, CH$_3$CN and SO$_2$ are also derived. Constrains on the density of the gas are estimated for two of the protostellar cores.} {We find that the dust condensations are in various evolutionary stages. The powerhouse of the cluster, mm1a, harbours a hot core with $T\\sim 220~(7510~M_ {\\sun})$. Two of the three identified outflows originate from the vicinity of mm1, which is probably the main powerhouse in the region. To zoom in on the innermost region around the protostars, and study the physical properties of the individual potentially star-forming cores, we carried out a comprehensive program to observe the region at high spatial resolution with the Plateau de Bure Interferometer\\footnote{IRAM is supported by INSU/CNRS (France), MPG (Germany) and IGN (Spain).} at 97~GHz and 241~GHz, and the Submillimeter Array\\footnote{The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica.} at 338~GHz. The new observations reach a resolution down to ~$0.6''$, corresponding to $\\sim 1100$~AU at the distance of the source. \\citet{05358-beuther} studied the continuum emission of this dataset. They identified four compact protostellar sources in the region; mm1 is resolved into two continuum peaks, mm1a and mm1b, with a projected linear separation of $\\sim 1700$~AU. A mid-infrared source \\citep{2006MNRAS.369.1196L}, and a compact 3.6~cm continuum source \\citep{05358-beuther} coincide with mm1a, which is also associated with the class II methanol masers detected by \\citet{2000A&A...362.1093M}. The previously identified source mm2 resolves into several sub-sources; however, only one of them (mm2a, according to the nomenclature used by \\citealt{05358-beuther}) is a protostellar source, while the others are probably caused by the outflows in the region. The third source mm3 remains a single compact core even at the highest spatial resolution. In Table~\\ref{cores}, we report the positions of the four sources identified in the continuum emission by \\citet{05358-beuther}. In this paper, we discuss the spectral line observations complementing the continuum data discussed by \\citet{05358-beuther}. In section \\S\\ref{observations}, the different observations are presented. In section \\S\\ref{obs-res}, we discuss our results, and analyse the extended emission of low excitation molecular transitions (\\S\\ref{extended}), as well as the molecular spectra at the positions of the dust condensations(\\S\\ref{proto}). Finally, in section \\S\\ref{analysis} we derive the physical parameters of the gas around the protostars from the analysis of their spectra. In the following sections, we use the term protostar for young massive stellar objects which are still accreting material from the surroundings, independently whether they already started burning hydrogen or not. \\begin{table} \\centering \\caption{Positions of the four dust condensations in IRAS\\,05358+3543 \\citep[from][]{05358-beuther}.\\label{cores}} \\begin{tabular}{lcc} \\multicolumn{1}{c}{Source} &\\multicolumn{1}{c}{R.A. [J2000]}&\\multicolumn{1}{c}{Dec. [J2000]}\\\\ \\hline \\hline mm1a&05:39:13.08&35:45:51.3\\\\ mm1b&05:39:13.13&35:45:50.8\\\\ mm2a&05:39:12.76&35:45:51.3\\\\ mm3&05:39:12.50&35:45:54.9\\\\ \\hline \\hline \\end{tabular} \\end{table} ", "conclusions": "Our new interferometric data resolve at least four cores in the high-mass protocluster IRAS\\,05358+3543. By analysing the molecular spectrum of each condensations, we characterised the properties of the gas surrounding it. Our main results are summarised in the following: \\begin{itemize} \\item the main powerhouse of the region, mm1a, harbours a hot core with $T\\sim 220$~K, and the central heating source has a chemical timescale of $10^{4.6}$~yr. Our data suggest that mm1a might host a massive circumstellar disk; \\item although the properties of the mm continuum emission of mm1b are very similar to the one of mm1a, the two sources differ significantly in the cm and mid-infrared spectrum. This core could be in an earlier stage of star formation than mm1a, since no molecular emission is detected toward it, with the only exception of the $5 \\to 4$~C$^{34}$S transition; \\item given the low abundance of methanol, mm2 could be a low--intermediate mass protostar; \\item strong emission is detected in several molecular species to the north-west of mm2, at a position where no continuum emission is detected. We suggest that this is caused by the interaction of the outflows with the ambient molecular cloud; \\item the least active source, mm3, could be a starless massive core, since it is cold ($T<20$~K), with a large reservoir of accreting material ($M\\sim 19~M_\\odot$), but no molecular emission peaks on it. \\end{itemize}" }, "0710/0710.5714_arXiv.txt": { "abstract": "Fusion reactions in the crust of an accreting neutron star are an important source of heat, and the depth at which these reactions occur is important for determining the temperature profile of the star. Fusion reactions depend strongly on the nuclear charge $Z$. Nuclei with $Z\\le 6$ can fuse at low densities in a liquid ocean. However, nuclei with $Z=8$ or 10 may not burn until higher densities where the crust is solid and electron capture has made the nuclei neutron rich. We calculate the $S$ factor for fusion reactions of neutron rich nuclei including $^{24}$O + $^{24}$O and $^{28}$Ne + $^{28}$Ne. We use a simple barrier penetration model. The $S$ factor could be further enhanced by dynamical effects involving the neutron rich skin. This possible enhancement in $S$ should be studied in the laboratory with neutron rich radioactive beams. We model the structure of the crust with molecular dynamics simulations. We find that the crust of accreting neutron stars may contain micro-crystals or regions of phase separation. Nevertheless, the screening factors that we determine for the enhancement of the rate of thermonuclear reactions are insensitive to these features. Finally, we calculate the rate of thermonuclear $^{24}$O + $^{24}$O fusion and find that $^{24}$O should burn at densities near $10^{11}$ g/cm$^3$. The energy released from this and similar reactions may be important for the temperature profile of the star. ", "introduction": "Nuclei accreting onto a neutron star undergo a variety of reactions. First at low densities, conventional thermonuclear fusion takes place, see for example \\cite{rpash}. Next as nuclei are buried to higher densities, the rising electron Fermi energy induces a series of electron captures \\cite{gupta}. Finally at very high densities, nuclei can fuse via pycnonuclear reactions. These reactions are induced by the quantum zero point motion \\cite{pycno}. The energy released, and the densities at which reactions occur, are important for determining the temperature profile of neutron star crusts. Superbursts are very energetic X-ray bursts from accreting neutron stars that are thought to involve the unstable thermonuclear burning of carbon \\cite{superbursts, superbursts2}. However, some simulations do not reproduce the conditions needed for carbon ignition because they have too low temperatures \\cite{superignition}. An additional heat source, from fusion or other reactions, could raise the temperature and allow carbon ignition at densities that reproduce observed burst frequencies. Recently the cooling of two neutron stars has been observed after extended outbursts \\cite{Wijnands, cackett}. These outbursts heated the crusts out of equilibrium and then the cooling time was measured as the crusts returned to equilibrium. The surface temperature of the neutron star in KS 1731-260 decreased with an exponential time scale of 325 $\\pm$ 100 days while MXB 1659-29 has a time scale of 505 $\\pm$ 59 days \\cite{cackett}. These cooling times depend on the thermal conductivity of the crust and the initial temperature profile. Comparing these observations of relatively rapid cooling to calculations by Rutledge et al. \\cite{rutledge} and Shternin et al. \\cite{shternin} suggests that the crust has a high thermal conductivity. However, if the initial temperature profile of the crust is peaked near the surface, then this peak could quickly diffuse to the surface and lead to rapid cooling. Therefore, cooling time scales are also sensitive to the initial temperature profile, and this depends on heating from nuclear reactions at moderate densities in the crust. Gupta et al. have calculated heating from electron capture reactions in the outer crust \\cite{gupta}. While they find more heating than previous works, they still find no more than 0.4 MeV per nucleon total heating from all of the electron captures on any mass number $A$ system. Haensel and Zdunik have calculated pycnonuclear fusion reactions at great densities in the inner crust \\cite{haensel}. However, if reactions occur deep in the inner crust, most of the heat may flow in to the core instead of out towards the surface. As a result, there may be a smaller impact on the temperature profile of the outer crust. A low crust thermal conductivity, for example from an amorphous solid, could help explain superburst ignition. This could better insulate the outer crust and allow higher carbon ignition temperatures. However, a low thermal conductivity appears to be directly contradicted by the observed short crust cooling times. Furthermore, our molecular dynamics simulations in ref. \\cite{horowitz} and further results we present in Section \\ref{MD} find a regular crystal structure, even when the system has a complex composition with many impurities. We do not find an amorphous phase. These results will be discussed further in a later publication. We conclude that the thermal conductivity of the crust is high. If the thermal conductivity is high, one may need additional heat sources, at moderate densities, in order to explain superburst ignition. Although Gupta et al. find additional heating from electron captures to excited nuclear states, simple nuclear structure properties may provide a natural limit to the total heating from electron captures \\cite{brownprivate}. Haensel and Zdunik \\cite{haensel,haensel2007} consider heating from pycnonuclear reactions using a simple one component plasma model. They find that fusion reactions may not take place until relatively high densities above $10^{12}$ g/cm$^3$. However, their use of a one component plasma could be a significant limitation. Fusion reactions depend strongly on the nuclear charge $Z$. Therefore, the reaction rate may be highest for the rare impurities that have the lowest $Z$, instead of for nuclei of average charge. In this paper, we go beyond Haensel and Zdunik and consider a full mixture of complex composition instead of assuming one average charge and mass. We focus on thermonuclear and pycnonuclear reactions at densities around $10^{11}$ g/cm$^3$. This is near the base of the outer crust. Heat released at this density could be important for superburst ignition and for crust cooling times. Nuclei at this density are expected to be neutron rich. Furthermore, the other nearby ions strongly screen the Coulomb barrier and greatly enhance the rate of thermonuclear reactions. We begin by describing the initial composition. This includes neutron rich light nuclei such as $^{24}$O and $^{28}$Ne. We calculate cross sections and $S$ factors for $^{24}$O + $^{24}$O and $^{28}$Ne + $^{28}$Ne fusion using a simple barrier penetration model. Note that the dynamics of the neutron rich skins of these nuclei can enhance the cross section over that predicted by our simple barrier penetration model. This is a very interesting and open nuclear structure question, see for example \\cite{subbarrier}. Next, we use classical molecular dynamics simulations to determine the structure of the crust and screening factors for the enhancement of thermonuclear reactions. There are many previous calculations of screening factors for the one component plasma \\cite{ocpscreening} and for binary ion mixtures, see for example \\cite{bimscreening}. However, we are not aware of any previous calculations for a crystal of a complex multicomponent composition. Finally, we calculate reaction rates and conclude that $^{24}$O is expected to fuse at densities near $10^{11}$ g/cm$^3$ while $^{28}$Ne should react at densities near $10^{12}$ g/cm$^3$. Heat from these reactions may be important for determining the temperature profile of accreting neutron stars. ", "conclusions": "\\label{summary} Fusion reactions in the crust of an accreting neutron star are an important source of heat, and the depth at which these reactions occur is important for determining the temperature profile of the star. Fusion reactions depend strongly on the nuclear charge $Z$. Nuclei with $Z\\le 6$ can fuse at low densities in a liquid ocean. However, nuclei with $Z=8$ or 10 may not burn until higher densities where the crust is solid and electron capture has made the nuclei neutron rich. In Section \\ref{crosssections} we calculated the $S$ factor for fusion reactions of neutron rich nuclei including $^{24}$O + $^{24}$O and $^{28}$Ne + $^{28}$Ne. We used a simple barrier penetration model. We find that $S$ for $^{24}$O+$^{24}$O is over eight orders of magnitude larger than that for $^{16}$O+$^{16}$O. The $S$ factor could be further enhanced by dynamical effects involving the neutron rich skin of $^{24}$O. For example, the skins of the two nuclei could deform to form a neck that would reduce the Coulomb barrier. This possible enhancement in $S$ should be studied in the laboratory with neutron rich radioactive beams. In Section \\ref{MD} we modeled the structure of the crust with molecular dynamics simulations. We find that the crust of accreting neutron stars may contain micro-crystals or regions of phase separation. Nevertheless, the screening factors that we determined for the enhancement of the rate of thermonuclear reactions are insensitive to these features. Finally, we calculated in Section \\ref{reaction rates} the rate of thermonuclear $^{24}$O + $^{24}$O fusion and find that $^{24}$O should burn at densities near $10^{11}$ g/cm$^3$. This is a lower density than some previous estimates. The 0.52 MeV per nucleon energy released may be important for the temperature profile of the star. In future work, we will use our molecular dynamics results to study other properties of the crust such as its thermal conductivity. In addition, we will use these MD results to calculate pycnonuclear reaction rates for the fusion of $^{28}$Ne and other heavier nuclei." }, "0710/0710.2628_arXiv.txt": { "abstract": "We observed four southern AXPs in 1999 near 1400 MHz with the Parkes 64-m radio telescope to search for periodic radio emission. No Fourier candidates were discovered in the initial analysis, but the recent radio activity observed for the AXP XTE J1810$-$197 has prompted us to revisit these data to search for single radio pulses and bursts. The data were searched for both persistent and bursting radio emission at a wide range of dispersion measures, but no detections of either kind were made. These results further weaken the proposed link between rotating radio transient sources and magnetars. However, continued radio searches of these and other AXPs at different epochs are warranted given the transient nature of the radio emission seen from XTE J1810$-$197, which until very recently was the only known radio-emitting AXP. ", "introduction": "The detection of pulsed radio emission from the anomalous X-ray pulsar (AXP) XTE J1810$-$197 in 2006 \\cite{crh+06}, and more recently from a second AXP, 1E 1547.0$-$5408 \\cite{crh+07}, has renewed interest in searching for radio emission from these objects. In both of these cases, the radio activity is believed to be connected to the X-ray variability of the sources and is transient in nature (or at least highly variable). Given this transient behavior and that both persistent periodic emission and single pulses were detected from both AXPs, renewed searches of archival radio search data of AXPs at different epochs may reveal previously undetected radio signals from these sources. ", "conclusions": "We found no convincing radio signals in either the folding or single pulse searches. The derived upper limits on the radio emission from our AXP targets are presented in Table \\ref{tbl-1}. These are the most stringent radio upper limits to date for these sources. The estimated 1400 MHz luminosity limits on the periodic radio emission ($\\approxlt 1$ mJy kpc$^{2}$) are 2-3 times lower than those established for XTE J1810$-$197 prior to outburst. However, it is still conceivable that weak radio pulses are being emitted, but that they are below our detection threshold. The luminosity limits presented here for the periodic emission are lower by about two orders of magnitude than the 1400 MHz luminosity of the periodic pulsed radio emission from XTE J1810$-$197 soon after the radio emission was first detected ($\\sim 80$ mJy kpc$^{2}$) \\citep{crh+06}. Thus we would expect to be able to easily detect comparably strong radio emission if it were beamed toward us. Our luminosity limits on single radio pulses from our sources range from 22 to 69 Jy kpc$^{2}$ in the most conservative case, which is below the 1400 MHz luminosity of $\\approxgt 100$ Jy kpc$^{2}$ derived from the pulse strengths reported for XTE J1810$-$197 in its radio discovery paper \\citep{crh+06}. Since single pulses were detected from almost every rotation of XTE J1810$-$197, it is likely that we would have detected a large number of comparable pulses during our observations if such pulses were beamed toward us. Our non-detection of single pulses further weakens the hypothesis that rotating radio transients (RRATs) \\citep{mlm+06} and magnetars are linked. This has been weakened by two other recent results. First, the X-ray detection of the RRAT J1819$-$1458 shows that its emission is more typical of middle-aged pulsars than it is of magnetars \\citep{rbg+06}. Second, the nearby, rotation-powered pulsar PSR B0656+54 would probably have been identified as an RRAT if it were farther away \\citep{wsr+06}. We conclude from our results that any periodic or bursting radio emission from the four target AXPs is either very weak (below our detection thresholds), not beamed toward us, or non-existent or sporadic at the epoch of observation. This last possibility is suggested by the connection between the X-ray and radio activity observed for the two known radio-emitting magnetars to date. Continued radio searches of AXPs are therefore warranted given the apparent transient nature of the radio emission. Further details of this work and a more complete discussion of the results are presented in a recent journal article \\cite{chk07}. \\begin{table} \\begin{tabular}{lcccc} \\hline & \\tablehead{1}{r}{b}{1E 1048.1$-$5937} & \\tablehead{1}{r}{b}{AX J1845$-$0258\\tablenote{AXP candidate only}} & \\tablehead{1}{r}{b}{1E 1841$-$045} & \\tablehead{1}{r}{b}{1RXS J170849.0$-$400910} \\\\ \\hline Spin period (s) & 6.45 & 6.97 & 11.77 & 11.00 \\\\ Ephemeris reference & \\cite{kgc+01} & \\cite{tkk+98}\\tablenote{No period derivative available} & \\cite{gvd99} & \\cite{gk02} \\\\ Galactic longitude, latitude (deg) & 288.26, $-$0.52 & 29.52, 0.07 & 27.39, $-$0.01 & 346.47, 0.03 \\\\ $T_{\\rm sky}$ (K)\\tablenote{1374 MHz sky temperature estimated from \\cite{hss+82} assuming a spectral index of $-2.6$} & 9.1 & 12.3 & 13.2 & 16.3 \\\\ Observation MJD & 51378 & 51391 & 51382 & 51379 \\\\ Observation date & 1999 Jul 19 & 1999 Aug 1 & 1999 Jul 23 & 1999 Jul 20 \\\\ $S_{1400}$ (mJy)\\tablenote{1400 MHz flux density limit on pulsed emission estimated using the modified radiometer equation and an assumed duty cycle of 2.7\\%} & $\\approxlt 0.02$ & $\\approxlt 0.02$ & $\\approxlt 0.02$ & $\\approxlt 0.02$ \\\\ $S_{1400}$ single (mJy)\\tablenote{Range of single-pulse 1400 MHz flux limits for pulse time-scales 0.25-240 ms} & $\\approxlt 875$-50 & $\\approxlt 975$-60 & $\\approxlt 1000$-60 & $\\approxlt 1085$-65 \\\\ Distance (kpc)\\tablenote{Taken from \\cite{bri+06}. Question marks indicate significant uncertainty in the value} & $\\sim 5$? & $\\sim 8$? & $\\sim 7$ & $\\sim 8$? \\\\ $L_{1400}$ (mJy kpc$^{2}$)\\tablenote{1400 MHz luminosity limit on pulsed emission, assuming a 1 sr beaming fraction} & $\\approxlt 0.5$ & $\\approxlt 1.3$ & $\\approxlt 1.0$ & $\\approxlt 1.3$ \\\\ $L_{1400}$ single (Jy kpc$^{2}$)\\tablenote{Range of 1400 MHz luminosity limits on single pulses for pulse time-scales 0.25-240 ms} & $\\approxlt 22$-1.3 & $\\approxlt 62$-3.7 & $\\approxlt 49$-2.9 & $\\approxlt 69$-4.1 \\\\ \\hline \\end{tabular} \\caption{Radio Search Parameters and Results} \\label{tbl-1} \\end{table}" }, "0710/0710.0288_arXiv.txt": { "abstract": "The channeling effect of low energy ions along the crystallographic axes and planes of NaI(Tl) crystals is discussed in the framework of corollary investigations on WIMP Dark Matter candidates. In fact, the modeling of this existing effect implies a more complex evaluation of the luminosity yield for low energy recoiling Na and I ions. In the present paper related phenomenological arguments are developed and possible implications are discussed at some extent. ", "introduction": "It is known that ions (and, thus, also recoiling nuclei) move in a crystal in a different way than in amorphous materials. In particular, ions moving (quasi-) parallel to crystallographic axes or planes feel the so-called ``channeling effect'' and show an anomalous deep penetration into the lattice of the crystal \\cite{chan,nel63,oth2}; see Fig. \\ref{fg:schema_chan}. For example, already on 1957, a penetration of $^{134}$Cs$^+$ ions into a Ge crystal was observed to a depth of about 1000 \\AA \\, \\cite{Bre56}, larger than that expected in the case the ions would cross amorphous Ge ($\\simeq 50$ \\AA). Afterwards, high intensities of H$^+$ ions at 75 keV transmitted through thick (3000-4000 \\AA) single-crystal gold films in the $<110>$ directions were detected \\cite{nel63}. Other examples for keV range ions have been shown in ref. \\cite{Ras05} where 3 keV P$^+$ ions moving into layers of 500 \\AA \\, of various crystals were studied. The channeling effect is also exploited in high energy Physics e.g. to extract high energy ions from a beam by means of bent crystals or to study diffractive Physics by analysing scattered ions along the beam direction (see e.g. ref. \\cite{Baur00}). Recently \\cite{droby2} it has been pointed out the possible role which this effect can play in the evaluation of the detected energy of recoiling nuclei in crystals, such as the NaI(Tl)\\footnote{For completeness, it is worth to note that luminescent response for channeling in NaI(Tl) was already studied in ref. \\cite{lunt} for MeV-range ions.}. \\begin{figure} [!t] \\centering \\includegraphics[width=9.0cm] {fig1.eps} \\caption{Simplified schema of the channeling effect in the NaI(Tl) lattice. The axial channeling occurs when the angle of the motion direction of an ion with the respect to the crystallographic axis is less than a characteristic angle, $\\Psi_c$, depicted there (see for details Sec. 2). Two examples for channeled and unchanneled ions are also shown (dashed lines).} \\label{fg:schema_chan} \\end{figure} In fact, the channeling effect can occur in crystalline materials due to correlated collisions of ions with target atoms. In particular, the ions through the open channels have ranges much larger than the maximum range they would have if their motion would be either in other directions or in amorphous materials. Moreover, when a low-energy ion goes into a channel, its energy losses are mainly due to the electronic contributions. This implies that a channeled ion transfers its energy mainly to electrons rather than to the nuclei in the lattice and, thus, its quenching factor (namely the ratio between the detected energy in keV electron equivalent [keVee] and the kinetic energy of the recoiling nucleus in keV) approaches the unity. It is worth to note that this fact can have a role in corollary analyses in the Dark Matter particle direct detection experiments, when WIMP (or WIMP-like) candidates are considered. In fact, since the routine calibrations of the detectors are usually performed by using $\\gamma$ sources (in order to avoid induced radioactivity in the materials), the quenching factor is a key quantity to derive the energy of the recoiling nucleus after an elastic scattering. Generally, for scintillation and ionization detectors this factor has been inferred so far by inducing tagged recoil nuclei through neutron elastic scatterings \\cite{Mis_neut}; however, as it will be discussed in Sec. 3, the usual analysis carried out on similar measurements does not allow to account for the channeled events. A list of similar values for various nuclei in different detectors can be found e.g. in ref. \\cite{RNC}. In particular, commonly in the interpretation of the dark matter direct detection results in terms of WIMP (or WIMP-like) candidates the quenching factors are assumed to be constant values without considering e.g. their energy dependence, the properties of each specific used detector and the experimental uncertainties. An exception was in the DAMA/NaI corollary model dependent analyses for WIMP (or WIMP-like) candidates \\cite{RNC,ijmd,epj06,ijma2} where at least some of the existing uncertainties on the $q_{Na}$ and $q_I $ values, measured with neutrons, were included. In this paper the possible impact of the channeling effect in NaI(Tl) crystals is discussed in a phenomenological framework and comparisons on some of the corollary analyses carried out in terms of WIMP (or WIMP-like) candidates \\cite{RNC,ijmd,epj06,ijma2}, on the basis of the 6.3 $\\sigma$ C.L. DAMA/NaI model independent evidence for particle Dark Matter in the galactic halo\\footnote{We remind that various possibilities for some of the many possible astrophysical, nuclear and particle Physics scenarios have have been analysed by DAMA itself both for some WIMP/WIMP-like candidates and for light bosons \\cite{RNC,ijmd,epj06,ijma2,ijma}, while other corollary analyses are also available in literature, such as e.g. refs. \\cite{Bo03,Bo04,Botdm,khlopov,Wei01,foot,Saib,droby1,droby2}. Many other scenarios can be considered as well.}, are given. ", "conclusions": "In this paper the channeling effect of recoiling nuclei induced by WIMP and WIMP-like elastic scatterings in NaI(Tl) crystals has been discussed. Its possible effect in a reasonably cautious modeling has been presented as applied to some given simplified scenarios in corollary quests for the candidate particle for the DAMA/NaI model independent evidence. This further shows the role of the existing uncertainties and of the correct description and modeling of all the involved processes as well as their possible impact in the investigation of the candidate particle. Some of them have already been addressed at some extent, such as the halo modeling \\cite{halo,RNC,ijmd}, the possible presence of non-thermalized components in the halo (e.g. caustics \\cite{siki} or SagDEG \\cite{epj06} contributions), the accounting for the electromagnetic contribution to the WIMP (or WIMP-like) expected energy distribution \\cite{ijma2}, candidates other than WIMPs (e.g. \\cite{ijma} and in literature), etc.. Obviously, many other arguments can be addressed as well both on DM candidate particles and on astrophysical, nuclear and particle physics aspects; for more see \\cite{RNC,ijmd,ijma,epj06,ijma2} and in literature. In particular, we remind that different astrophysical, nuclear and particle Physics scenarios as well as the experimental and theoretical associated uncertainties leave very large space also e.g. for significantly lower cross sections and larger masses." }, "0710/0710.0846_arXiv.txt": { "abstract": "s{ The Alpha Magnetic Spectrometer (AMS), to be installed on the International Space Station (ISS) in 2008, is a cosmic ray detector with several subsystems, one of which is a proximity focusing Ring Imaging \\CK\\ (RICH) detector. This detector will be equipped with a dual radiator (aerogel+NaF), a lateral conical mirror and a detection plane made of 680 photomultipliers and light guides, enabling precise measurements of particle electric charge and velocity. Combining velocity measurements with data on particle rigidity from the AMS Tracker it is possible to obtain a measurement for particle mass, allowing the separation of isotopes. \\\\ A Monte Carlo simulation of the RICH detector, based on realistic properties measured at ion beam tests, was performed to evaluate isotope separation capabilities. Results for three elements --- H (Z=1), He (Z=2) and Be (Z=4) --- are presented. } \\vspace{-0.9cm} ", "introduction": " ", "conclusions": "AMS will provide a major improvement on existing data for isotopic abundances in cosmic rays. Simulation results indicate that the separation of light isotopes using the combination of RICH data and tracker rigidity measurements is feasible. The dual radiator configuration of NaF and aerogel makes isotope separation of light elements possible for energies in the range from 0.5 to 10 GeV/nucleon, approximately. Best mass resolutions are $\\sim$~2\\% at 3 GeV/nucleon for aerogel, and $\\sim$~3\\% at 1 GeV/nucleon for NaF. Techniques presented here may also be applied in the separation of antimatter isotopes which is of great importance in dark matter studies. \\vspace{-0.2cm}" }, "0710/0710.2558_arXiv.txt": { "abstract": "We highlight the potential importance of gaseous TiO and VO opacity on the highly irradiated close-in giant planets. The atmospheres of these planets naturally fall into two classes that are somewhat analogous to the M- and L-type dwarfs. Those that are warm enough to have appreciable opacity due to TiO and VO gases we term the ``pM Class'' planets, and those that are cooler, such that Ti and V are predominantly in solid condensates, we term ``pL Class'' planets. The optical spectra of pL Class planets are dominated by neutral atomic Na and K absorption. We calculate model atmospheres for these planets, including pressure-temperature profiles, spectra, and characteristic radiative time constants. Planets that have temperature inversions (hot stratospheres) of $\\sim$2000 K and appear ``anomalously'' bright in the mid infrared at secondary eclipse, as was recently found for planets \\hh\\ and \\hd, we term the pM Class. Molecular bands of TiO, VO, H$_2$O, and CO will be seen in emission, rather than absorption. This class of planets absorbs incident flux and emits thermal flux from high in their atmospheres. Consequently, they will have large day/night temperature contrasts and negligible phase shifts between orbital phase and thermal emission light curves, because radiative timescales are much shorter than possible dynamical timescales. The pL Class planets absorb incident flux deeper in the atmosphere where atmospheric dynamics will more readily redistribute absorbed energy. This leads to cooler day sides, warmer night sides, and larger phase shifts in thermal emission light curves. We briefly examine the transit radii for both classes of planets. The boundary between these classes is particularly dependent on the incident flux from the parent star, and less so on the temperature of the planet's internal adiabat (which depends on mass and age), and surface gravity. Around a Sun-like primary, for solar composition, this boundary likely occurs at $\\sim$0.04-0.05 AU, but uncertainties remain. We apply these results to pM Class transiting planets that are observable with the \\emph{Spitzer Space Telescope}, including \\hd, WASP-1b, TrES-3b, TrES-4b, \\hh, and others. The eccentric transiting planets HD 147506b and HD 17156b alternate between the classes during their orbits. Thermal emission in the optical from pM Class planets is significant red-ward of 400 nm, making these planets attractive targets for optical detection via Kepler, COROT, and from the ground. The difference in the observed day/night contrast between $\\upsilon$ Andromeda b (pM Class) and \\he\\ (pL Class) is naturally explained in this scenario. ", "introduction": "The blanket term ``hot Jupiter\" or even the additional term ``very hot Jupiter'' belies the diversity of these highly irradiated planets. Each planet likely has its own unique atmosphere, interior structure, and accretion history. The relative amounts of refractory and volatile compounds in a planet will reflect the parent star abundances, nebula temperature, total disk mass, location of the planet's formation within the disk, duration of its formation, and its subsequent migration (if any). This accretion history will give rise to differences in core masses, total heavy elements abundances, and atmospheric abundance ratios. Given this incredible complexity, it is worthwhile to first look for physical processes that may be common to groups of planets. In addition to a mass and radius, one can further characterize a planet by studying its atmosphere. The visible atmosphere is a window into the composition of a planet and contains clues to its formation history \\cp[e.g.,][]{Marley07b}. Of premier importance in this class of highly irradiated planets is how stellar insolation affects the atmosphere, as this irradiation directly affects the atmospheric structure, temperatures, and chemistry, the planet's cooling and contraction history, and even its stability against evaporation. Since irradiation is perhaps the most important factor in determining the atmospheric properties of these planets, we examine the insolation levels of the 23 known transiting planets. We restrict ourselves to those planets more massive than Saturn, and hence for now exclude treatment of the ``hot Neptune'' GJ 436b, which is by far the coolest known transiting planet. \\mbox{Figure~\\ref{flux}} illustrates the stellar flux incident upon the planets as a function of both planet mass (\\mbox{Figure~\\ref{flux}}\\emph{a}) and planet surface gravity (\\mbox{Figure~\\ref{flux}}\\emph{b}). In these plots diamonds indicate transiting planets and triangles indicate other interesting hot Jupiters, for which \\emph{Spitzer Space Telescope} data exist, but which do not transit. The first known transiting planet, \\hd, is seen to be fairly representative of these planets in terms of incident flux. Planets OGLE-TR-56b and OGLE-TR-132b are somewhat separate from the rest of the group because they receive the highest stellar irradiation. Both orbit their parent stars in less than 2 days and are prototypes of what has been called the class of ``very hot Jupiters'' \\cp{Konacki03,Bouchy04} with orbital periods less than 3 days. However, orbital period is a poor discriminator between ``very hot'' and merely ``hot,'' as \\he\\ clearly shows. Labeled a ``very hot Jupiter'' upon its discovery, due to its short 2.2 day period \\citep{Bouchy05}, \\he\\ actually receives a comparatively modest amount of irradiation due to its relatively cool parent star. Therefore, perhaps a classification based on incident flux, equilibrium temperature, or other attributes would be more appropriate. In this paper we argue that based on the examination of few physical processes that two classes of hot Jupiter atmospheres emerge with dramatically different spectra and day/night contrasts. Equilibrium chemistry, the depth to which incident flux will penetrate into a planet's atmosphere, and the radiative time constant as a function of pressure and temperature in the atmosphere all naturally define two classes these irradiated planets. Our work naturally builds on the previous work of \\ct{Hubeny03} who first investigated the effects of TiO and VO opacity on close-in giant planet atmospheres as a function of stellar irradiation. These authors computed optical and near infrared spectra of models with and without TiO/VO opacity. In general they found that models with TiO/VO opacity feature temperature inversions and molecular bands are seen in emission, rather than absorption. Two key questions from the initial \\ct{Hubeny03} investigation were addressed but could not be definitely answered were: 1) if a relatively cold planetary interior would lead to Ti/V condensing out deep in the atmosphere regardless of incident flux, thereby removing gaseous TiO and VO, and, 2) if this condensation did not occur, at what irradiation level would TiO/VO indeed be lost at the lower atmospheric temperatures found at smaller incident fluxes. Later \\ct{Fortney06} investigated model atmospheres of planet \\hh\\ including TiO/VO opacity at various metallicities. Particular attention was paid to the temperature of the deep atmosphere \\emph{P-T} profiles (as derived from an evolution model) in relation to the Ti/V condensation boundary. Similar to \\ct{Hubeny03}, they found a temperature inversion due to absorption by TiO/VO and computed near and mid-infrared spectra that featured emission bands. Using the \\emph{Spitzer} InfraRed Array Camera (IRAC) \\ct{Harrington07} observed \\hh\\ in secondary eclipse with \\emph{Spitzer} at 8 $\\mu$m and derived a planet-to-star flux ratio consistent with a \\ct{Fortney06} model with a temperature inversion due to TiO/VO opacity. At that point, looking at the work of \\ct{Fortney06} and especially \\ct{Hubeny03}, \\ct{Harrington07} could have postulated that all objects more irradiated than \\hh\\ may possess inversions due to TiO/VO opacity, but given the single-band detection of \\hh, caution was in order. More recently, based on the four-band detection of flux from \\hd\\ by \\ct{Knutson08}, \\ct{Burrows07c} find that a temperature inversion, potentialy due to TiO/VO opacity, is necessary to explain this planet's mid-infrared photometric data. Based on their new \\hd\\ model and the previous modeling investigations these authors posit that planets warmer than \\hd\\ may features inversions, while less irradiated objects such as \\he\\ do not, and discuss that photochemical products and gaseous TiO/VO are potential absorbers which may lead to this dichotomy. We find, as has been previously shown, that those planets that are warmer than required for condensation of titanium (Ti) and vanadium (V)-bearing compounds will possess a temperature inversion at low pressure due to absorption of incident flux by TiO and VO, and will appear ``anomalously'' bright in secondary eclipse at mid-infrared wavelengths. Thermal emission in the optical will be significant \\cp{Hubeny03,Lopez07}. Furthermore, here we propose that these planets will have large day/night effective temperature contrasts. We will term these very hot Jupiters the ``pM Class,'' meaning gaseous TiO and VO are the prominent absorbers of optical flux. The predictions of equilibrium chemistry for these atmospheres are similar to dM stars, where absorption by TiO, VO, H$_2$O, and CO is prominent \\cp{Lodders02b}. Planets with temperatures below the condensation curve of Ti and V bearing compounds will have a gradually smaller mixing ratio of TiO and VO, leaving Na and K as the major optical opacity sources \\cp*{BMS}, along with H$_2$O, and CO. We will term these planets the pL class, similar to the dL class of ultracool dwarfs. These planets will have relatively smaller secondary eclipse depths in the mid infrared and significantly smaller day/night effective temperature contrasts. As discussed below, published \\emph{Spitzer} data are consistent with this picture. The boundary between these classes, at irradiation levels (and atmospheric temperatures) where Ti and V may be partially condensed is not yet well defined. In this paper we begin by discussing the observations to date. We then give an overview of our modeling methods and the predicted chemistry of Ti and V. We calculate pressure-temperature (\\emph{P-T}) profiles and spectra for models planets. For these model atmospheres we then analyze in detail the deposition of incident stellar flux and the emission of thermal flux, and go on to calculate characteristic radiative time constants for these atmospheres. We briefly examine transmission spectra before we apply our models to known highly irradiated giant planets. Before our discussion and conclusions we address issues of planetary classification. ", "conclusions": "Though 1D radiative-convective equilibrium model atmospheres we have addressed the class of atmospheres, the pM Class, for which TiO and VO are extremely strong visible absorbers \\cp{Hubeny03}. This absorbed incident flux drives these planets to have hot ($\\sim$2000+ K) stratospheres. Therefore, these planets will appear very bright in the mid-infrared, with brightness temperatures larger than their equilibrium temperatures. This is the case for \\hh\\ \\cp{Harrington07}, and \\hd\\ \\ct{Knutson08}. In addition, these planets will have large day/night temperature contrasts because radiative time constants at photospheric pressures are much shorter than reasonable advective timescales. The hottest point of the planet should be the substellar point, which absorbs the most flux, leading to perhaps negligible phase shift between the times of maximum measured thermal emission and when the day side is fully visible. This appears to be the situation for planet $\\upsilon$ And b, observed by \\ct{Harrington06}. Given that its irradiation level is intermediate between \\hd\\ and \\hh\\ we find that this planet is pM Class. Due to the fast radiative times, the day side of these planets may have \\emph{P-T} profiles that do not deviate much from radiative equilibrium models. Although atmospheric dynamics will surely be vigorous, winds will be unable to advect gas before it cools to space. For these planets, high irradiation, the presence of gaseous TiO and VO, a hot stratosphere, the location of the hottest atmosphere at the substellar point, and a large day/night temperature contrast all go hand-in-hand. On the other hand, we have shown that in the pL Class, dominated by absorption by H$_2$O, Na, and K, photospheric pressures and temperatures prevail such that advective timescales and radiative timescales are similar \\cp[see also][]{Seager05}. Since atmospheric dynamics will be important for the redistribution of energy, the consequences for the structure and thermal emission of these atmospheres will be quite complex. The efficiency of energy redistribution will vary with planetary irradiation level, surface gravity, and rotation rate. pL Class planets will have smaller day/night temperature contrasts and measurable phase shifts in thermal emission light curves that will be wavelength-dependent. In addition, these planets may show variability in secondary eclipse depth \\cp[e.g.~][]{Rauscher07}. However, without a better understanding of the dynamics it is difficult to make detailed predictions at this time. Secondary eclipse depths should range somewhere between values expected for a ``full redistribution'' model and inefficient redistribution. The published secondary eclipse data for pL Class planets \\T\\ and \\he\\ are all consistent with this prediction \\cp{Fortney05,Fortney07b}. In addition, the 8 $\\mu$m light curves for 51 Peg b and \\hd\\ \\cp{Cowan07} and \\he\\ \\cp{Knutson07b} are consistent with this prediction as well. Examination of \\mbox{Figure~\\ref{flux}} shows that HD 179949b is a pM Class planet, and indeed \\ct{Cowan07} found the largest phase variation it their small sample for this planet, but the unknown orbital inclination makes definitive conclusions difficult. Additional observational results will soon help to test the models presented here. We find that transiting planets WASP-1b, TrES-4b, TrES-3b, OGLE-Tr-10b, and TrES-2b will be in the pM Class, along with non-transiters $\\upsilon$ And b and HD 179949. The low-irradiation boundary of this class is not yet clear, and planet \\hd\\ shows that temperature inversions persist to irradiation levels where TiO/VO are expected to begin being lost to condensation \\cp{Burrows07c}. At still lower irradiation levels, the limited data for \\he\\ lead us to conclude it is pL Class \\cp{Fortney07b}. Secondary eclipse data for XO-2b, HAT-P-1b, and WASP-2b will be important is determining how temperature inverstions (and the TiO/VO abundances) wane with irradiation level. Just as in dM and dL stars, the condensation of Ti and V is expected to be gradual process, so we fully expect transition objects between the distinct pM and pL class members. We will soon have additional information that will help shed light on the atmosphere of \\hd\\. The 8 $\\mu$m light curve for \\hd\\ obtained by \\ct{Cowan07} shows little phase variation. However, if our theory connecting TiO/VO opacity and temperature inversions to large day/night contrasts is correct, we expect to see a large variation. Soon H.~Knutson and collaborators will obtain half-orbit light curves for \\hd\\ at 8 and 24 $\\mu$m. The quality should be comparable to that obtained by \\citet{Knutson07b} for \\he, and will put our theory to the test. HD 147506, with an eccentricity of 0.517 \\cp{Bakos07b}, and HD 17156, with an eccentricity of 0.67 \\cp{Fischer07,Barbieri07}, will be extremely interesting cases as the flux they receive varies by factors of 9 and 26, respectively, between periapse and apoapse. They should each spend part of their orbits as pL class and part as pM class. This makes predictions difficult, but large day/night temperatures differences at apoapse are likely. There has recently been considerable discussion on the relative merits of multi-dimensional dynamical models and 1D radiative-convective model atmospheres for these highly irradiated planets. Both kinds of studies provide interesting predictions. While it could be claimed that 1D radiative-convective models are unrealistic because they ``lack dynamics,'' they do include very detailed chemistry, vast opacity databases, and advanced non-gray radiative transfer, which all dynamics models for these planets lack. Given the large diversity in predictions among the various dynamical models, which use a host of simplifications, the next step will be combining dynamics and radiative transfer, an idea which has been mentioned or advanced by a number of authors \\cp{Seager05,Barman05,Fortney06b,Burrows06,Dobbs07}. Eventually we will be able to give up 1D $f$-type parameters to treat the incident flux in a more realistic fashion for these exotic atmospheres. We recently started working toward this goal in \\ct[][see also \\citealp{Dobbs07}]{Fortney06b}; work continues, and we believe it will have a promising future. We think that the predictions we have made here with a 1D model will provide a framework for understanding the observations to come. Additional observations for both pL and pM class planets, along with additional theoretical and modeling efforts, should further clarify our understanding." }, "0710/0710.3148_arXiv.txt": { "abstract": "In this paper, warm inflationary models on a brane scenario are studied. Here we consider slow-roll inflation and high-dissipation regime in a high-energy scenario. General conditions required for these models to be realizable are derived. We describe scalar and tensor perturbations for these scenarios. Specifically we study power-law potentials considering a dissipation parameter to be a constant on the one hand and $\\phi$ dependent on the other hand. We use recent astronomical observations to restrict the parameters appearing in our model. ", "introduction": "It is well known that many long-standing problems of the Big Bang model, namely the horizon problem, flatness, homogeneity and the numerical density of monopoles, may find a natural explanation in the frame of the inflationary universe model \\cite{Guth1981,Inflation,Inflationary}. Perhaps the most relevant feature of the inflationary universe model is that it provides a causal interpretation for the origin of the observed anisotropy in the cosmic microwave background (CMB) radiation, and also the distribution of large-scale structures \\cite{WMAP1,WMAP3}. But the inflationary universe model has problems too. One of the problems in it is how to attach the observed universe to the end of the inflationary epoch; there are three possible solutions to this problem: reheating \\cite{KolbandTurner}, preheating \\cite{TBKLS} and warm inflation \\cite{Berera1995,deOliveira1998,Berera1999,Bellini1998}. In this work we focus on the latter. In standard inflationary universe models, the acceleration of the universe is driven by a scalar field (named inflaton) with a specific scalar potential. These kinds of model are divided into two regimes, the slow-roll and reheating ephocs. In the slow-roll period the universe inflates and all interactions between the inflaton scalar field and any other field are typically neglected. Subsequently, a reheating period is invoked to end the period of inflation. After reheating, the universe is filled with radiation \\cite{afterReheating,Inflation}, and then the universe gets connected with the Big Bang model. Warm inflation is an alternative mechanism to have successful inflation and avoid the reheating period \\cite{Berera1995}. In this kind of model, dissipative effects are important during inflation, so that radiation production occurs concurrently with the inflationary expansion. The inflaton interacts with a thermal bath via a friction term, where phenomenologically the decay of the scalar field is described by means of an interaction Lagrangian. For instance, the authors of Ref.\\cite{new} take the interaction terms of the form $\\frac{1}{2}\\lambda^2\\phi^2\\chi^2$ and $g\\chi\\bar{\\psi}\\psi$, where the inflationary period presents a two-stage decay chain $\\phi\\rightarrow\\chi\\rightarrow\\psi$. In this case, they reported that the damping term $\\Gamma$ becomes $\\frac{\\lambda^3g^2\\phi}{256\\pi^2}$. Note that if the scalar field changes a little bit, then the friction coefficient $\\Gamma$ remains almost constant. From the point of view of statistical mechanics, the interaction between quantum fields and a thermal bath could be illustrated by a general fluctuation-dissipation relation \\cite{new2}. Warm inflation was criticized on the basis that the inflaton cannot decay during the slow-roll phase \\cite{new3}. However, in recent years, it has been shown that the inflaton can indeed decay during the slow-roll phase (see \\cite{new4} and references therein) whereby it now rests on solid theoretical grounds. On the other hand, the inclusion of the damping term into the model needs a very special scheme: for instance, in Ref.\\cite{Profe} it was found that when $\\Gamma$ is constant, in the slow-roll approximation a wrong value for the number of e-fold was obtained: meanwhile, in the power-law approach this problem is absent. Warm inflation ends when the universe heats up to become radiation dominated. At this epoch the universe stops inflating and `smoothly' enters into a radiation dominated Big Bang phase \\cite{Berera1995}. The matter components of the universe are created by the decay of either the remaining inflationary field or the dominant radiation field \\cite{TaylorandBerera2000}. In standard inflationary universe models the quantum fluctuations associated to the inflaton scalar field generate the density perturbations seeding the structure formation at a late time in the evolution of the universe. Instead, in warm inflation models, the density fluctuations arise from thermal rather than quantum fluctuations \\cite{Berera2004,Berera1995}. These fluctuations have their origin in the hot radiation and influence the inflaton through a friction term in the equation of motion of the inflaton scalar field \\cite{Berera1996}. Several aspects of warm inflationary universe models have been studied in the past few years. The goal of the present work is to investigate warm inflationary models on brane scenarios, where the total energy density $\\rho=\\rho_{\\phi}+\\rho_{\\gamma}$ is found on the brane \\cite{Profe2}. The universe is filled with a self-interacting scalar field of energy density $\\rho_{\\phi}$ and a radiation field with energy density $\\rho_{\\gamma}$. The motivation for introducing brane scenarios is the increasing interest in higher dimensional cosmological models, motivated by superstring theory, where the matter fields (related to open string modes) are confined to a lower dimensional brane, while gravity (closed string modes) can propagate in the bulk \\cite{Strings}. Shiromizu et al. \\cite{Shiromizu} have found the four-dimensional Einstein's equations projected onto the brane. These projections introduce some differences in the fundamental field equations, such as the Friedmann equation and, therefore, in the equations that describes the linear perturbations theory \\cite{Brane}. Our aim is quantify the modifications of the warm inflation model in the brane scenario for arbitrary inflaton potentials on the brane. In order to do this we study the linear theory of cosmological perturbations for a warm inflationary model on a brane. The perturbations are expressed in term of different parameters appearing in our model: these parameters will be constrained from the WMAP three-year data \\cite{WMAP3}. In section II we describe the dynamics of our model and we establish some approximations that we use in this work. In section III we investigate the linear theory of perturbations. Here we calculate the scalar perturbations in the longitudinal gauge and also the tensor perturbations. In section IV we take a power-law potential and we investigate the high-energy and high-dissipation regime considering a power-law dissipation coefficient $\\Gamma$. Finally, in section V, we give some conclusions. We have used units in which $c=\\hbar=1$. ", "conclusions": "In this paper we have considered a warm inflationary scenario on a brane. We have restricted ourselves to a high-dissipation and high-energy regime. In the slow-roll approximation we have found a general relationship between radiation and scalar field densities; see Eq.(\\ref{eq21}). In relation to the perturbations we have considered that the there does no exist any interaction between the brane and the bulk, i.e., we have neglected back-reaction due to metric perturbations in the fifth dimension. We note that a full investigation is required to discover when back-reaction will have a significant effect in the perturbations. With the above-mentioned restriction we have obtained explicitly the contributions of the adiabatic and entropy modes. We have shown that the dissipation parameter plays a crucial role in producing the entropy mode (see Eq.(\\ref{eq62b})). A general relation for the density perturbations is given in Eq.(\\ref{eq41}). The tensor pertubations are generated via stimulated emission into the existing thermal background (see Eq.(\\ref{eq52})) and the tensor-scalar ratio is modified by a temperature-dependent factor. We have studied a power-law potential for different dependence of the dissipation coefficient $\\Gamma$. From the normalization of the WMAP three-year data, the potential becomes of the order of $V(\\phi_0)\\sim10^{-20}M_4^4$ when it leaves the horizon at the scale of $k_0=0.002\\text{Mpc}^{-1}$. As in the situation studied in Ref.\\cite{Chaotic}, the value of the potential depends on $M_5$. Here we have considered $M_5\\sim10^{-5}M_4$. To fulfill the approximations in our model we have restricted the range of the parameters in which warm inflation on a brane can occur (see FIGS. \\ref{fig:HT}, \\ref{fig:VM5}). In order to show some explicit results we have chosen the following set of values for the parameters, $T_r\\simeq T=3.5\\times10^{-6}M_4$, $r=1000$ and $s=100$. First, we have considered a chaotic potential and a constant dissipation coefficient; for this case we have found that the spectrum is driven toward an scale-invariant spectrum and the running of the spectral index is extremely small. The situation was similar when we considered a variable dissipation coefficient. In the case of a power-law potential with $n=4$ and a constant dissipation coefficient we have found that the value to $n_s$ is smaller than the previous case but out of the range given by WMAP three-year data. On the other hand, the running of the spectral index is closer to the value given by WMAP three-year data but one order of magnitud smaller. Finally the best situation is found when we consider a constant dissipation coefficient and a $n=6$ potential. In this case both parameters, $n_s$ and $\\alpha_s$, are in the ranges specified by WMAP. It is necessary to note that with our approximations we have found that for the case in which $n=6$ and $m=0$ is favored in the light of the recent results reported by WMAP three-year data. On the other hand, we have considered some approximations; as a result, we were not able to find any real solution in the allowed range of parameters. For example, in the slow-roll approximation, in the case $n=4$ and, when we try to consider a variable coefficient of dissipation $\\Gamma$, we did not find any real solution. However, we think that we could find a solution if in place of using this approximation we assume a power-law for the scale factor. We intend to return to this point (and others) in the near future." }, "0710/0710.1651_arXiv.txt": { "abstract": "We report on an investigation of the environments of the SLACS sample of gravitational lenses. The local and global environments of the lenses are characterized using SDSS photometry and, when available, spectroscopy. We find that the lens systems that are best modelled with steeper than isothermal density profiles are more likely to have close companions than lenses with shallower than isothermal profiles. This suggests that the profile steepening may be caused by interactions with a companion galaxy as indicated by N-body simulations of group galaxies. The global environments of the SLACS lenses are typical of non-lensing SDSS galaxies with comparable properties to the lenses, and the richnesses of the lens groups are not as strongly correlated with the lens density profiles as the local environments. Furthermore, we investigate the possibility of line-of-sight contamination affecting the lens models but do not find a significant over-density of sources compared to lines of sight without lenses. ", "introduction": "The Sloan Lens ACS Survey \\citep[SLACS;][]{bolton} is a large sample of strong gravitational lenses derived from the Sloan Digital Sky Survey (SDSS). The lens sample has proved to be particularly useful due to the high quality of the data; all of the SLACS lenses have known lens and source redshifts, stellar velocity dispersions for the lensing galaxies have been measured from the SDSS spectra, all of the systems have {\\em Hubble Space Telescope} ({\\em HST}) ACS imaging in two bands for accurate non-parametric lens modelling, and all of the sources are extended and therefore provide additional constraints on the mass profile of the lensing galaxy. Furthermore, the density of background sources in the {\\em HST} imaging allows a weak lensing analysis of the ensemble sample of lenses \\citep{gavazzi}; the major drawback of the lens sample is that it is unlikely to find any variable sources that would provide time delays. \\citet{koopmansSLACS} have used the SLACS lenses to show that early-type galaxies have isothermal total inner density profiles with very little intrinsic scatter (approximately 6 per cent) assuming a uniformity in the environments of the lenses that has not been rigorously tested. While the SLACS lenses seem to lie on the Fundamental Plane \\citep{bolton,treu,bolton07} and do not differ noticeably from other SDSS galaxies with similar luminosities and stellar velocity dispersions, the local environments of the lenses might affect the mass profiles of the lensing galaxies \\citep[e.g.,][]{rusin,dobke,augerb}. If some of the lenses are being perturbed by neighbouring galaxies the intrinsic scatter of the density slope for {\\em isolated} early-type galaxies might be even smaller than 6 per cent. Furthermore, it has been suggested that line-of-sight (LOS) contamination significantly affects the SLACS lenses \\citep{guimaraes} and the density slope might be expected to be shallower than originally reported. We report on a spectroscopic and photometric evaluation of the environments and lines of sight of the 15 SLACS lenses investigated by \\citet{koopmansSLACS}. A weighting scheme is used to determine the effective number of potential perturbing companions to each lens galaxy. We also characterize the `richness' of the global environment of each lens field and quantify the number of galaxies along the LOS to the lens. Throughout this paper the term `global environment' is used to describe the group, cluster, or field in which the lens resides while the `local environment' describes the environment within $\\approx 100$ \\hinv kpc of the lensing galaxy. A $\\Lambda$CDM cosmology with $\\Omega_M = 0.27$ and $\\Omega_\\Lambda = 0.73$ is used to determine all physical distances, which are measured in \\hinv units. ", "conclusions": "We find that the global environments of the SLACS lenses are typical of other massive early-type galaxies found in the SDSS. Two of the steeper than isothermal systems lie in very over-dense regions but the remaining lenses all have richness values that fall near the peak of the richness distribution (Figure \\ref{figure_comp_rich}). Furthermore, the suggestion that the SLACS lenses are affected by LOS contamination does not seem to be merited by the data. While there are other galaxies along the lines of sight to the lens systems, the LOS densities do not significantly deviate from the densities along comparable lines of sight. We therefore expect that the parameter estimates should not be affected as proposed by \\citet{guimaraes}. The SDSS photometric data indicate that the SLACS lenses are only slightly more likely to be associated with companion galaxies than comparable lenses selected from the SDSS, though the uncertainty of the photometric identifications makes the difference negligible. We therefore conclude that SLACS lenses do lie in typical environments both globally and locally. However, we also find that lens systems with steeper than isothermal density slopes are preferentially associated with companion galaxies compared to lenses with shallower density slopes. N-body simulations suggest that interactions with neighbouring galaxies can induce a steepening in the density slope \\citep{dobke} and there are other lens systems with companion galaxies that are found to be best modelled with steeper than isothermal profiles \\citep[e.g.,][]{rusin,augerb}. The interaction-induced steepening is a transient effect and the density profile of the galaxy will return to isothermal approximately 0.5-2 Gyr after the encounter with the neighbour \\citep{dobke}. This may account for the large range of $N_w$ for lenses with nearly isothermal profiles; isothermal lenses with a companion may be in the relaxed state before or after an encounter with the neighbour galaxy. This stripping mechanism may also account for local observations of dark matter deficient galaxies \\citep[e.g.,][]{romanowsky,proctor}. We note that one steeper-profile system, SDSSJ1250+0523, is not photometrically associated with any neighbouring galaxies; this perhaps illustrates the limits of using photometry to find perturbing companions or demonstrates that other factors also influence the slope of the density profile. Additionally, we have not found an environmental bias to account for the shallower lenses. However, if a lens galaxy is embedded in a cluster, the joint profile of the cluster and galaxy would tend to be modelled with a shallower than isothermal profile if a single-component power law is used (ignoring interactions between the two halos). This effect is dependent on the location of the lens with respect to the centre of the cluster, which our photometric analysis is unable to address. More complete field spectroscopy would better characterize the global environments of these lens systems and allow correlations between the mass slopes and cluster centre offsets to be investigated. Furthermore, a spectroscopic investigation of the local environments of the complete sample of SLACS early-type lenses would confirm the correlation indicated by our photometric analysis and provide strong evidence for truncation caused by galaxy interactions." }, "0710/0710.0711_arXiv.txt": { "abstract": "{ We present results of $H\\alpha$ imaging for 42 galaxies in the nearby low-density cloud Canes Venatici I populated mainly by late-type objects. Estimates of the $H\\alpha$ flux and integrated star formation rate ($SFR$) are now available for all 78 known members of this scattered system, spanning a large range in luminosity, surface brightness, $HI$ content and $SFR$. Distributions of the CVnI galaxies versus their $SFR$, blue absolute magnitude and total hydrogen mass are given in comparison with those for a population of the nearby virialized group around M81. We found no essential correlation between star formation activity in a galaxy and its density environment. A bulk of CVnI galaxies had enough time to generate their baryon mass with the observed $SFR$. Most of them possess also a supply of gas sufficient to maintain their observed $SFR$s during the next Hubble time. } ", "introduction": "The distribution over the sky of 500 galaxies of the Local volume with distances within 10 Mpc shows considerable inhomogeneities due to the presence of groups and voids. Apart from several virialized groups, like the group around the galaxy M81, an amorphous association of nearby galaxies in the Canes Venatici constellation was noted by many authors (Karachentsev 1966, de Vaucouleurs 1975, Vennik 1984, Tully 1988). The boundaries of it are rather uncertain. Roughly, in a circle of radius $\\sim20\\degr$ around the galaxy NGC~4736 there are about 80 known galaxies with $D<10$ Mpc, which corresponds to a density contrast on the sky $\\Delta N/N\\sim4$. Inside of this complex the distribution of galaxies is also inhomogeneous, showing some clumps differing in their location on the sky and distances in depth. About 70\\% of the population of the cloud accounts for irregular dwarf galaxies, whose masses are obviously insufficient to keep such a system in the state of virial equilibrium. As data on distances of galaxies accumulated, a possibility appeared of studying the kinematics of the cloud in details. As was shown by Karachentsev et al. (2003), the CVnI cloud is in a state close to the free Hubble expansion, having a characteristic crossing time of about 15 Gyr. Being a scattered system with rare interactions between galaxies, the nearby CVnI cloud is a unique laboratory for studying star formation processes in galaxies running independently, without a noticeable external influence. Kennicutt et al. (1989), Hoopes et al. (1999), van Zee (2000), Gil de Paz et al. (2003), James et al. (2004) and Hunter \\& Elmegreen (2004) conducted observations in the $H\\alpha$ line of three dozen galaxies of this complex, which made it possible to determine the star formation rate ($SFR$) in them. However, more than half of other members of the cloud proved to be out of vision of these authors. Our task consisted in the completion of $H\\alpha$-survey of the population of the CVnI cloud. The results of our observations and their primary analysis are presented in this paper. ", "conclusions": "A systematic survey of $H{\\alpha}$-emission in the nearest scattered cloud CVnI shows that in most of its galaxies the process of active star formation is on despite the low density contrast of this cloud and rather rare interaction between its members. By making full use of our $H{\\alpha}$-survey, we can estimate the mean density of $SFR$ in the cloud. A conical volume in the distance range from 2 to 10 Mpc, resting in a sky region of $\\sim1500$ square degrees, makes 153 Mpc$^3$. In this volume we have a summary value $\\Sigma(SFR)=18.6M_{\\sun}$yr$^{-1}$, which yields an average density of star formation rate $\\dot{\\rho}_{SFR}=0.12M_{\\sun}$yr$^{-1}$Mpc$^{-3}$. The obtained value turns out to be a little less than $\\dot{\\rho}_{SFR}=0.165M_{\\sun}$yr$^{-1}$Mpc$^{-3}$ for a ``sell of homogeneity'' embracing the group M81 (Karachentsev \\& Kaisin, 2007). According to Nakamura et al. (2004), Martin et al. (2005) and Hanish et al. (2006), the average star formation rate per 1 Mpc$^3$ at the present epoch (z=0) is (0.02--0.03)$M_{\\sun}$yr$^{-1}$Mpc$^{-3}$. Consequently, the CVnI cloud has excess of $SFR$ density 4--6 times higher in comparison with the mean global quantity, which roughly corresponds to the cloud density contrast on the sky, $\\Delta N/N\\sim4$. Herefrom we conclude that the CVnI cloud is characterised by a usual norm of star formation in its galaxies." }, "0710/0710.3120.txt": { "abstract": "We present a new measurement of the volumetric rate of Type Ia supernova up to a redshift of 1.7, using the Hubble Space Telescope (HST) GOODS data combined with an additional HST dataset covering the North GOODS field collected in 2004. We employ a novel technique that does not require spectroscopic data for identifying Type Ia supernovae (although spectroscopic measurements of redshifts are used for over half the sample); instead we employ a Bayesian approach using only photometric data to calculate the probability that an object is a Type Ia supernova. This Bayesian technique can easily be modified to incorporate improved priors on supernova properties, and it is well-suited for future high-statistics supernovae searches in which spectroscopic follow up of all candidates will be impractical. Here, the method is validated on both ground- and space-based supernova data having some spectroscopic follow up. We combine our volumetric rate measurements with low redshift supernova data, and fit to a number of possible models for the evolution of the Type Ia supernova rate as a function of redshift. The data do not distinguish between a flat rate at redshift $>$ 0.5 and a previously proposed model, in which the Type Ia rate peaks at redshift $\\sim$ 1 due to a significant delay from star-formation to the supernova explosion. Except for the highest redshifts, where the signal to noise ratio is generally too low to apply this technique, this approach yields smaller or comparable uncertainties than previous work. ", "introduction": "%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5 The empirical evidence for the existence of dark energy came from observations of Type Ia supernovae~\\citep{bib:riess, bib:P99, bib:P03}, which are believed to arise from the thermonuclear explosion of a progenitor white dwarf after it approaches the Chandrasekhar mass limit~\\citep{bib:chan}. However, the physics of Type Ia supernova production is not well understood. The two most plausible scenarios for the white dwarf to accrete the necessary mass are the single degenerate case, where the white dwarf is located in a binary system; and the double degenerate case, where two white dwarfs merge. The Type Ia supernova rate is correlated with the star formation history (SFH), and thus a measurement of the rate as a function of redshift helps constrain the possible type Ia progenitor models. In addition to its importance for understanding Type Ia supernovae as astronomical objects, a good grasp of the Type Ia supernova rate to high redshifts is important for the next generation of proposed space-based supernova cosmology experiments, such as SNAP~\\citep{bib:snap}. It is therefore of great practical interest to determine the rate of Type Ia supernovae at redshifts $>$ 1. The subject of Type Ia supernova rates has been addressed by many authors in the past. Existing rate measurements have been mostly limited to redshift ranges $<$ 1: the results of~\\cite{bib:cappellaro},~\\cite{bib:hardin},~\\cite{bib:madgwick}, and~\\cite{bib:blanc} measure the rates at redshifts $\\leq$ $\\sim$0.1;~\\cite{bib:neill},~\\cite{bib:tonry}, and~\\cite{bib:pain}, at intermediate redshifts of 0.47, 0.50, and 0.55, respectively; and~\\cite{bib:barris}, up to a redshift of 0.75. The only published measurement of the rates at redshifts $>$ 1 is that of~\\cite{bib:dahlen}, who analyzed the GOODS dataset. There are several important differences that distinguish our work from that of~\\cite{bib:dahlen}. First, we augment the GOODS sample with the HST data collected during the Spring-Summer 2004 high redshift supernova searches. Second, our methods of calculating the control time (the time during which a supernova search is potentially capable of finding supernova candidates) and the efficiency to identify a supernova are based on a detailed Monte Carlo simulation technique using a library of supernova templates. Third, we adopt a novel approach to typing supernovae, using photometric data and a Bayesian probability method described in~\\cite{bib:ourpaper}. The Bayesian technique is able to perform classification using only photometric data, and therefore does not require spectroscopic follow up. Optionally, photometric or spectroscopic redshifts can be used to improve the classification accuracy. Our initial requirements on potential supernova candidates are more stringent in terms of the number of points on the light curve and the signal to noise of those points than those of~\\cite{bib:dahlen}; thus some of the candidates they identified will fail our cuts. However, we are able to reliably separate Type Ia supernova from other supernovae types based on their Bayesian probability, with an efficiency that is readily quantifiable, thus allowing us to use larger data samples. Our approach therefore avoids the problems that arise in estimating the efficiency for the decision to schedule spectroscopic follow up based on a potentially low signal-to-noise initial detection. The Bayesian classification technique uses photometric data, and does not require any spectroscopic followup. This is an advantage for future large-area surveys (such as the Dark Energy Survey, Pan-STARRS, and LSST) that will discover thousands of supernova candidates, but are unlikely to be able to obtain spectroscopic data for all of them, to distinguish Type Ia supernovae from core collapse supernovae and other variable objects. The technique described here can be considered a prototype of the kind of analysis that could be performed on these future large data sets to identify Type Ia supernovae for cosmological studies. There is a clear trade-off involved in using photometric measurements alone: if the quality of the photometric data is poor, then the efficiency of this technique to identify Type Ia supernovae is reduced; on the other hand, this technique enables larger samples of Type Ia from imaging surveys to be identified for cosmological studies, without the need for time-consuming spectroscopic follow up. Note that although the method is able to perform the supernova typing with photometric data alone (\\emph{i.e.}, it does not require spectroscopic data, either redshifts or types), it is certainly able to use the extra information that is available, and in fact 70\\% of the supernova candidates discussed in the present work have redshifts which were obtained spectroscopically. It is also worth noting that while in this paper we only analyze the Type Ia supernova rates, the Bayesian classification technique can be used to classify other types as well, making it possible to measure the rates of non-Type Ia supernovae in a similar fashion. These analyses will be presented in future publications. The paper is organized as follows. In section~\\ref{sec:data} we describe the data samples used in the analysis. In section~\\ref{sec:candselection} we describe the supernova candidate selection and typing process. In section 4 we calculate the control time, survey area, and search efficiency, and determine the volumetric Type Ia supernova rate from our data sample. A comparison of the rates with those reported in the literature is given in section~\\ref{sec:comp}, and fits of the rates to different models relating the Type Ia supernova rates to the SFH are given in section~\\ref{sec:sf}. A summary is given in section~\\ref{sec:concl}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{sec:concl} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% We have analyzed the rates of Type Ia supernovae up to a redshift of 1.7 using two samples collected with the HST: the GOODS data, and the 2004 ACS sample collected in the Spring-Summer 2004 covering the GOODS North field. Using only the data from two broadband filters, F775W and F850LP, we applied a novel technique for identifying Type Ia supernovae based on a Bayesian probability approach. This method allows us to automatically type supernova candidates in large samples, properly taking into account all known sources of systematic error. We also make use of the best currently available full spectral templates for five different supernova types for the candidate typing, as well as for calculating the efficiency of our supernova search, and the control time. These templates will undoubtedly be improved over the next several years as more supernova data becomes available. Current and upcoming supernova surveys will not only provide a better understanding of individual supernova types, but may also uncover new types of supernovae, which can then be added to the Bayesian classification framework. Likewise, a better understanding of the many parameters that affect supernova observations will improve the classification scheme, which will result in better constraints on the measured rates. The calculations of the supernova finding efficiency, the control time, and the survey area are all done taking into account the specific observing configurations pertinent for the surveys, such as exposure times, cadences, and the orientations of the GOODS tiles. We carried out a comparison of the predicted and observed numbers of supernovae in redshift bins of $\\Delta \\bar{z}$ = 0.1, for two different models of the Type Ia supernova rates: a redshift-independent rate and a power-law redshift-dependent rate. We find that the available data fit both models equally well. For comparison with previous work, particularly that of~\\cite{bib:dahlen}, who also analyzed a large subset of the data used here, we calculated the volumetric Type Ia supernova rates in four redshift bins, 0.2 $\\leq$ $\\bar z$ $<$ 0.6, 0.6 $\\leq$ $\\bar z$ $<$ 1.0, 1.0 $\\leq$ $\\bar z$ $<$ 1.4, and 1.4 $\\leq$ $\\bar z$ $<$ 1.7. We find that our results are generally consistent with those of~\\cite{bib:dahlen}. Due to the larger of number supernova candidates which this Bayesian classification technique makes available, we obtain smaller or equal uncertainties in all the bins up to $z$ = 1.7. In the highest redshift bin we obtain a larger uncertainty because the signal to noise ratio is generally too low to apply this technique. We fitted the resulting rates to two leading models used in recent literature: the two-component model and a Gaussian time delay model. The former model implies an increase in the Type Ia supernova rates at highest redshifts; while the latter, a decrease. We find that the statistics of the present sample does not definitively discriminate between the two scenarios -- only one supernova in this work and two supernovae in~\\cite{bib:dahlen} contribute to the important highest-redshift bin. Significantly larger surveillance time would be required to arrive at a conclusive statement on the trends for the Type Ia rates at high redshifts. In the future, several ambitious new surveys are planned that will collect photometric data for thousands of supernovae in order to improve the constraints on dark energy. Individual spectroscopic follow up for every supernova candidate is likely to be impractical in these surveys. The Bayesian classification method described here has the ability to classify supernovae using photometric measurements alone, and is a promising technique for these future surveys. %%%%%%%%" }, "0710/0710.4129.txt": { "abstract": "{} {To compute the chemical evolution of spiral bulges hosting Seyfert nuclei, based on updated chemical and spectro-photometrical evolution models for the bulge of our Galaxy, to make predictions about other quantities measured in Seyferts, and to model the photometric features of local bulges. The chemical evolution model contains updated and detailed calculations of the Galactic potential and of the feedback from the central supermassive black hole, and the spectro-photometric model covers a wide range of stellar ages and metallicities.} {We computed the evolution of bulges in the mass range $2\\times 10^{9}-10^{11}M_{\\odot}$ by scaling the efficiency of star formation and the bulge scalelength as in the inverse-wind scenario for elliptical galaxies, and considering an Eddington limited accretion onto the central supermassive black hole.} {We successfully reproduced the observed relation between the mass of the black hole and that of the host bulge. The observed nuclear bolometric luminosity emitted by the supermassive black hole is reproduced only at high redshift or for the most massive bulges; in the other cases, at $z \\simeq 0$ a rejuvenation mechanism is necessary. The energy provided by the black hole is in most cases not significant in the triggering of the galactic wind. The observed high star formation rates and metal overabundances are easily achieved, as well as the constancy of chemical abundances with the redshift and the bulge present-day colours. Those results are not affected if we vary the index of the stellar IMF from $x=0.95$ to $x=1.35$; a steeper IMF is instead required in order to reproduce the colour-magnitude relation and the present $K$-band luminosity of the bulge.} {We show that the chemical evolution of the host bulge, with a short formation timescale of $\\sim 0.1$ Gyr, a rather high efficiency of star formation ranging from $11$ to $50$ Gyr$^{-1}$ according to the bulge mass and an IMF flatter with respect to the solar neighbourhood, combined with the accretion onto the black hole is sufficient to explain the main observed features of Seyfert galaxies.} ", "introduction": "The outstanding question of the co-evolution of Active Galactic Nuclei (AGNs) and their host galaxies has received considerable attention in the past decades, since various pieces of evidence pointed to a link between the formation of supermassive black holes (BHs) and the formation and evolution of their host spheroids: for example, the usual presence of massive dark objects at the centre of nearby spheroids (Ford et al. 1997; Ho 1999; Wandel 1999); the correlation between the BH mass and the stellar velocity dispersion of the host (for quiescent galaxies, Ferrarese \\& Merritt 2000; Gebhardt et al. 2000a; Tremaine et al. 2002; for active galaxies, Gebhardt et al. 2000b; Ferrarese et al. 2001; Shields et al., 2003; Onken et al. 2004; Nelson et al. 2004) or its mass (Kormendy \\& Richstone 1995; Magorrian et al. 1998; Marconi \\& Hunt 2003; Dunlop et al. 2003); the similarity between light evolution of quasar (QSO) population and the star formation history of galaxies (Cavaliere \\& Vittorini 1998; Haiman et al. 2004); the establishment of a good match among the optical QSO luminosity function, the luminosity function of star-forming galaxies and the mass function of dark matter halos (DMHs) at $z\\sim 3$ (Haenhelt et al., 1998). The most widely accepted explanation for the luminosity emitted by an AGN, is radiatively efficient gas accretion onto a central supermassive BH. %Some theoretical model link quasar activity with major merger events %in a hierarchical galaxy formation process (Kauffmann \\& Haenhelt %2000; Kauffmann et al., 2003). The outflows from AGNs can profoundly affect the evolution of the host galaxy, e.g. by quenching or inducing the star formation (e.g., see Ciotti \\& Ostriker 2007, and references therein). The mutual feedback between galaxies and QSOs was used as a key to solve the shortcomings of the semianalytic models in galaxy evolution, e.g. the failure to account for the surface density of high-redshift massive galaxies (Blain et al., 2002; Cimatti et al., 2002) and for the $\\alpha$-enhancement as a function of mass (Thomas et al., 2002), since it could provide a way to invert the hierarchical scenario for the assembly of galaxies and star formation (see e.g. Monaco et al., 2000; Granato et al., 2004; Scannapieco et al., 2005). The study of the chemical abundances of the QSOs was first undertaken by Hamann \\& Ferland (1993), who combined chemical evolution and spectral synthesis models to interpret the N~{\\scshape v}/C~{\\scshape iv} and N~{\\scshape v}/He~{\\scshape ii} broad emission line ratios, and found out that the high metallicities and the abundance ratios of the broad-line region are consistent with the outcomes of the models for giant elliptical galaxies (Arimoto \\& Yoshii, 1987; Matteucci \\& Tornamb\\`e, 1987; Angeletti \\& Giannone, 1990), where the timescales of star formation and enrichment are very short and the initial mass function (IMF) is top-heavy. In the same year, Padovani \\& Matteucci (1993) and Matteucci \\& Padovani (1993) employed the chemical evolution model of Matteucci (1992) to model the evolution of radio-loud QSOs, which are hosted by massive ellipticals, following in detail the evolution of several chemical species in the gas. They supposed that the mass loss from dying stars after the galactic wind provides the fuel for the central BH and modeled the bolometric luminosity as $L_{bol}=\\eta\\dot{M}c^2$, with a typical value for the efficiency of $\\eta = 0.1$ and were successful in obtaining the estimated QSO luminosities and the observed ratio of AGN to host galaxy luminosity. Then, they studied the evolution of the chemical composition of the gas lost by stars in elliptical galaxies and spiral bulges for various elements (C, N, O, Ne, Mg, Si and Fe), and found out that due to the high star-formation rate (SFR) of spheroids at early times the standard QSO emission lines were naturally explained. The relatively weak observed time dependence of the QSO abundances for $t \\gtrsim 1$ Gyr was also predicted. The model of Matteucci \\& Padovani (1993) still followed the classic wind scenario, where the efficiency of star formation decreases with increasing galactic mass and which was found to be inconsistent with the correlation between spheroid mass and $\\alpha$-enhancement (Matteucci 1994). Moreover, Padovani \\& Matteucci (1993) pointed out that if all mass lost by stars in the host galaxy after the wind were accreted by the central BH, the final BH mass would be up to two orders of magnitude larger than observed. Other works (Fria\\c ca \\& Terlevich 1998; Romano et al. 2002; Granato et al., 2004), which had a more refined treatment of gas dynamics, limited their analysis of chemical abundances to the metallicity $Z$ and the [Mg/Fe] ratio and their correlation with the galactic mass. All these studies were mainly devoted to studying the co-evolution of radio-loud QSOs and their host spheroids, which are elliptical galaxies. Now we want to extend the approach of Padovani \\& Matteucci (1993) to AGNs hosted by spiral bulges, with a more recent chemical evolution model for the bulge with the introduction of the treatment of feedback from the central BH and a more sophisticated dealing of the accretion rate. Since Seyfert nuclei are preferentially hosted by disk-dominated galaxies (Adams, 1977; Yee, 1983; MacKenty, 1990; Ho et al. 1997) our study can be applied to this class of objects. The paper is organized as follows: in \\S 2 we illustrate the chemical and photometrical evolution model, in \\S 3 we show our calculations of the potential energy and of the feedback from supernovae (SNe) and from the AGN, in \\S 4 we discuss our results concerning the black hole masses and luminosities, the chemical abundances and the photometry, and in \\S 5 we draw some conclusions. ", "conclusions": "" }, "0710/0710.5628_arXiv.txt": { "abstract": "Dust has been detected in the recurrent nova RS\\,Ophiuchi on several occassions. I model the historical mid-infrared photometry and a recent Spitzer Space Telescope spectrum taken only half a year after the 2006 eruption. The dust envelope is little affected by the eruptions. I show evidence that the eruptions and possibly the red giant wind of RS\\,Oph may sculpt the interstellar medium, and show similar evidence for the recurrent dwarf nova T\\,Pyxidis. ", "introduction": "\\subsection{Asymptotic Giant Branch stars and red supergiants} Stars with an initial mass between $\\sim1$ and $\\sim40$ M$_\\odot$ become hydrogen/helium-shell-burning Asymptotic Giant Branch (AGB) stars or core-helium-burning red supergiants (RSG). These phases are characterised by cool, molecular atmospheres giving rise to M spectral types (S or C for some chemically peculiar AGB stars), strong radial pulsations of this atmosphere on timescales of a year or more, and a high luminosity ($L>2,000$ L$_\\odot$). The pulsation may lift the atmosphere high enough such that dust can condense \\citep{BowenWillson1991}. Radiation pressure on the grains then drives a dust wind which, via collisions with molecular hydrogen, drags the gas along with it. At a typical wind speed $v_\\infty\\sim5$ to 30 km s$^{-1}$ these stars lose mass at rates from $\\dot{M}\\sim10^{-7}$ to over $10^{-4}$ M$_\\odot$ yr$^{-1}$ \\citep{vanLoonEtal1999}. In AGB stars this leads to the removal of the mantle and the premature death of the star, leaving the truncated core behind as a cooling white dwarf. The more massive RSGs either explode or evolve back to higher surface temperatures, and it is not yet clear how much mass they will have shed during the preceding RSG stage. Mass-loss rates of cool, luminous stars are most easily estimated from the reprocessed radiation emitted by the dust grains at infrared (IR) wavelengths, which is particularly conspicuous in the $\\lambda\\sim10$ to 40 $\\mu$m region. The main problem with this method is that the dust constitutes only a minor fraction of the total mass in the wind, typically $\\sim1$:$200$ \\citep{Knapp1985} but poorly known for all but the dustiest stars. Carbon monoxide has strong rotational transitions at mm wavelengths which can be detected in relatively nearby stars, but despite being the most abundant molecule after H$_2$ this too is a trace species of which the mass fraction is uncertain --- and modelling the CO line requires knowledge of the temperature profile throughout the envelope which depends, amongst other things, on the dust content. \\subsection{First ascent Red Giant Branch stars} Low-mass stars ($M_{\\rm initial}<2$ M$_\\odot$) evolve along a first ascent Red Giant Branch (RGB) as hydrogen-shell burning stars with an inert helium core. They reach a maximum luminosity of only $L_{\\rm tip}\\sim2,000$ L$_\\odot$, and it becomes problematic for them to drive a wind. The situation is worsened by the fact that many RGB stars do not pulsate strongly and slowly enough to sufficiently increase the scaleheight, and the dust condensation is therefore unlikely to reach completeness. With a low, uncertain dust:gas mass ratio and possibly higher dilution of the dust envelope as it kinematically decouples from the bulk gas, measured mass-loss rates will tend to be too low. The threshold below which this happens is not well known, partly because of uncertainties in the opacity of the grains as they nucleate and grow: \\citet{GailSedlmayr1987} estimate $\\dot{M}\\gg 10^{-6}$ M$_\\odot$ yr$^{-1}$ but \\citet{NetzerElitzur1993} place it at $\\dot{M}>10^{-7}$ M$_\\odot$ yr$^{-1}$. \\citet{JudgeStencel1991} show that RGB mass loss must be driven by another mechanism, but that this appears to be similarly efficient as a dust-driven wind. The warmer stars further down the RGB have more prominent chromospheres, which may be accompanied by the generation of Alfv\\'en or acoustic waves that provide a possible alternative mechanism for driving a wind. Mass-loss rates estimated from the IR emission for the brightest, dustiest RGB stars are typically $\\dot{M}\\sim10^{-7}$ to $10^{-6}$ M$_\\odot$ \\citep{vanLoonEtal2006,OrigliaEtal2007}. Mass-loss rates from $\\dot{M}\\sim10^{-9}$ to $10^{-6}$ M$_\\odot$ yr$^{-1}$ have been determined from the blue-displaced cores of strong optical absorption lines and emission in some of these lines, but these are very uncertain too because of the sensitivity to the precise excitation and ionization conditions in the wind. For the most luminous RGB stars the IR and optical estimates tend to agree within an order of magnitude \\citep{McDonaldvanLoon2007}. \\begin{figure}[!t] \\plotone{Jacco_vanLoon_fig1.eps} \\caption{RS\\,Oph compared to isochrones from \\citet{GirardiEtal2000}.} \\end{figure} ", "conclusions": "Dust in RS\\,Oph has been detected in between several eruptions, and at only half a year since the 2006 eruption. The silicate dust species and mass-loss rate of $\\dot{M}\\sim2$ to $3\\times10^{-8}$ M$_\\odot$ yr$^{-1}$ are not at odds with the expectations for well-developed dust-driven winds of single stars. But because RS\\,Oph has a relatively warm photosphere, weak pulsation and not very high luminosity, this might in fact suggest that the mass-loss rate {\\it is} enhanced by the effects of the companion white dwarf. The eruptions do {\\it not} seem to have a dramatic effect on the dust envelope. An accretion rate onto the white dwarf surface of $\\dot{M}>10^{-7}$ M$_\\odot$ yr$^{-1}$ would probably require an enhanced flow through the Lagrangian point L$_1$. Considering the pattern in the time intervals between historic eruptions, including the 1907 eruption \\citep{Schaefer2004}, one might heuristically expect the next eruption to be due around 2015 (Fig.\\ 10). \\begin{figure}[!b] \\plotone{Jacco_vanLoon_fig10.eps} \\caption{Recent outbursts of RS\\,Oph and a heuristic prediction for the next.} \\end{figure}" }, "0710/0710.2839_arXiv.txt": { "abstract": "{% Neutron stars contain persistent, ordered magnetic fields that are the strongest known in the Universe. However, their magnetic fluxes are similar to those in magnetic A and B stars and white dwarfs, suggesting that flux conservation during gravitational collapse may play an important role in establishing the field, although it might also be modified substantially by early convection, differential rotation, and magnetic instabilities. The equilibrium field configuration, established within hours (at most) of the formation of the star, is likely to be roughly axisymmetric, involving both poloidal and toroidal components. The stable stratification of the neutron star matter (due to its radial composition gradient) probably plays a crucial role in holding this magnetic structure inside the star. The field can evolve on long time scales by processes that overcome the stable stratification, such as weak interactions changing the relative abundances and ambipolar diffusion of charged particles with respect to neutrons. These processes become more effective for stronger magnetic fields, thus naturally explaining the magnetic energy dissipation expected in magnetars, at the same time as the longer-lived, weaker fields in classical and millisecond pulsars.} ", "introduction": "The purpose of this talk is to present and discuss some of the physical processes that are likely to be relevant in determining the structure and evolution of magnetic fields in neutron stars, almost regardless of their (so far largely unknown) internal composition and state of matter. For this reason, I do not discuss the fascinating, exotic issues of quark matter, superfluidity, superconductivity, and the like, but emphasize the much more pedestrian concepts of stable stratification and non-ideal magnetohydrodynamics (MHD) processes such as ambipolar diffusion. A review with a very different focus has recently been given by Geppert (2006). ", "conclusions": "The extremely strong magnetic fields found in neutron stars are nevertheless weak enough to be balanced by small perturbations in the density, pressure, and chemical composition of the stably stratified, degenerate, multi-species fluid found in their interior. The field likely originates from the fairly strong magnetic flux inherited from the progenitor star, possibly modified by a combination of convection, differential rotation, and magnetic instabilities acting during the short, protoneutron star stage following collapse. The equilibrium field set up during this stage likely involves linked toroidal and poloidal field components. It can decay through dissipative processes such as weak interactions and ambipolar diffusion, which change the chemical composition of the matter, allowing it to move. These processes become particularly effective at high field strengths, possibly accounting for the energy release inside magnetars, which also keeps the stellar interior hot enough for the magnetic field to rearrange substantially. This forces changes as well in the crust of the star, which can be broken by strong fields, whereas weaker ones can evolve by a combination of Hall drift and resistive dissipation. The internal dissipation processes are ineffective for pulsar-strength fields, which can easily survive for a pulsar lifetime. None of these processes depends essentially on the exotic properties of a neutron star, such as Cooper pairing or quark matter, which can however modify the time scales for these processes to occur." }, "0710/0710.5244_arXiv.txt": { "abstract": "In this paper we present a deep and homogeneous i-band selected multi-waveband catalogue in the COSMOS field covering an area of about $0.7\\sq\\degr$. Our catalogue with a formal 50\\% completeness limit for point sources of $i\\sim 26.7$ comprises about 290~000 galaxies with information in 8 passbands. We combine publicly available u, B, V, r, i, z, and K data with proprietary imaging in H band. We discuss in detail the observations, the data reduction, and the photometric properties of the H-band data. We estimate photometric redshifts for all the galaxies in the catalogue. A comparison with 162 spectroscopic redshifts in the redshift range $ 0 \\lsim z \\lsim 3$ shows that the achieved accuracy of the photometric redshifts is \\mbox{$\\Delta z / (z_{spec}+1) \\lsim 0.035$} with only $\\sim 2$\\% outliers. We derive absolute UV magnitudes and investigate the evolution of the luminosity function evaluated in the restframe UV (1500~\\AA). { There is a good agreement between the LFs derived here and the LFs derived in the FORS Deep Field. We see a similar brightening of M$^\\ast$ and a decrease of $\\phi^\\ast$ with redshift.} The catalogue including the photometric redshift information is made publicly available. ", "introduction": "\\label{sec:intro} In the last decade our knowledge about the evolution of global galaxy properties over a large redshift range has improved considerably. The 2dF Galaxy Redshift Survey (2dFGRS; \\citealt{colless:1}), the Sloan Digital Sky Survey (SDSS; \\citealt{stoughton:1}), and the 2MASS survey \\citep{2MASS} have provided very large local galaxy samples with spectroscopic and/or photometric information in various passbands. Thanks to these data sets we are now able to assess very accurate local ($z\\sim 0.1$) reference points for many galaxy evolution measurements like the luminosity function, the star formation activity, the spatial clustering of galaxies, the stellar population, the morphology, etc. In the redshift range between $0.2 \\lsim z \\lsim 1 $ pioneering work has been done in the context of the Canada France Redshift Survey \\citep{lilly:3}, the Autofib survey \\citep{ellis:1} and in the Canadian Network for Observational Cosmology survey \\citep{yee:1}. They provide accurate distances and absolute luminosities by spectroscopic followup of optically selected galaxies, thus being able to probe basic properties of galaxy evolution. Moreover the K20-survey \\citep{cimatti:2} as well as the MUNICS survey \\citep{drory:2,feulner:1} extend the analysis into the near infrared regime (for $0.2 \\lsim z \\lsim 1.5 $). An important step towards probing the galaxy properties also in the high redshift regime around $z \\sim 3$ and $z\\sim 4$ was the work of \\citet{steidel_lbg:1} and \\citet{steidel_lbg:2}. They used colour selection to discriminate between low and high redshift galaxies \\citep[see ][for a review]{giavalisco:3}. The so-called Lyman-break galaxies (LBGs, mainly starburst galaxies at high redshift) are selected by means of important features in the UV spectrum of star-forming galaxies. The next milestones in pushing the limiting magnitude for detectable galaxies to fainter and fainter limits were the space based Hubble Deep Field North (HDFN; \\citealt{HDF96}) and Hubble Deep Field South \\citep[HDFS; ][]{HDFS00,HDFS00a} \\citep[see ][for a review]{ferguson:1}. Although of a limited field of view of about $5\\sq\\arcmin$ only, the depth of the HDFs allowed the detection of galaxies up to a redshift of 5 and even beyond. In the past years the space based HDFs were supplemented by many more multi-band photometric surveys like the NTT SUSI deep Field (NDF; \\citealt{arnouts_ntt}), the Chandra Deep Field South (CDFS; \\citealt{arnouts_cdfs}), the William Herschel Deep Field (WHDF; \\citealt{mccracken:1,metcalf:1}), the Subaru Deep Field/Survey (SDF; \\citealt{maihara:1,ouchi:2}), the COMBO-17 survey \\citep{combo17:1}, FIRES \\citep{labbe:1}, the FORS Deep Field (FDF; \\citealt{fdf_data}), the Great Observatories Origins Deep Survey (GOODS; \\citealt{giavalisco:1}), the Ultra Deep Field (UDF and UDF-Parallel ACS fields; \\citealt{giavalisco:2,bunker:1, bouwens:2004}), the VIRMOS deep survey \\citep{lefevre:3}, GEMS \\citep{rix:1}, the Keck Deep Fields \\citep{sawicki:2}, and the Multiwavelength Survey by Yale-Chile (MUSYC; \\citealt{gawiser:1, quadri:1}). With the advent of all these deep multi-band photometric surveys the photometric redshift technique (essentially a generalisation of the drop-out technique) can be used to identify high-redshift galaxies. Photometric redshifts are often determined by means of template matching algorithm that applies Bayesian statistics and uses semi-empirical template spectra matched to broad-band photometry (see also \\citealt{baum:1, koo:1, brunner:1, Soto:1, benitez:1, bender:1, borgne:1, firth:1}). Redshifts of galaxies that are several magnitudes fainter than typical spectroscopic limits can be determined reliably with an accuracy of \\mbox{$\\Delta z / (z_{spec}+1)$} of 0.02 to 0.1. In this context the COSMOS survey (\\citet{scoville:1}; see also http://www.astro.caltech.edu/cosmos/ for an overview) combines deep to very-deep multi-waveband information in order to extend the analysis of deep pencil beam surveys to a much bigger volume, thus being able to drastically increase the statistics and detect also very rare objects. For this, the survey covers an area of about $2\\sq\\degr$ with imaging by space-based telescopes (Hubble, Spitzer, GALEX, XMM, Chandra) as well as large ground based telescopes (Subaru, VLA, ESO-VLT, UKIRT, NOAO, CFHT, and others). In this paper we combine publicly available u, B, V, r, i, z, and K COSMOS data with proprietary imaging in the H band to derive a homogeneous multi-waveband catalogue suitable for deriving accurate photometric redshifts. In Section~\\ref{sec:imaging_data} we give an overview of the near-infrared (NIR) data acquisition and we describe our 2-pass data reduction pipeline used to derive optimally (in terms of signal-to-noise for faint sources) stacked images in Section~\\ref{sec:imaging_reduction}. We also present NIR galaxy number counts and compare them with the literature.\\\\ In Section~\\ref{sec:isel_catalogue} we present the deep multi-waveband i-band selected catalogue and discuss its properties, whereas the data reduction of the spectroscopic redshifts is described in Section~\\ref{sec:spec}. In Section~\\ref{sec:photoz} we present the photometric redshift catalogue, discuss the accuracy of the latter and show the redshift distribution of the galaxies. In Section~\\ref{sec:uvlf} we derive the redshift evolution of the restframe UV luminosity function and luminosity density at 1500~\\AA\\ from our i-selected catalogue before we summarise our findings in Section~\\ref{sec:summary_conclusion}. We use AB magnitudes and adopt a $\\Lambda$ cosmology throughout the paper with \\mbox{$\\Omega_M=0.3$}, \\mbox{$\\Omega_\\Lambda=0.7$}, and \\mbox{$H_0=70 \\, \\mathrm{km} \\, \\mathrm{s}^{-1} \\, \\mathrm{Mpc}^{-1}$}. ", "conclusions": "\\label{sec:summary_conclusion} In this paper we present the data acquisition and reduction of NIR Js, H, and K' bands in the COSMOS field. We describe a 2-pass reduction pipeline to reduce NIR data. The 2-pass pipeline is optimised to avoid flat-field errors introduced if the latter are constructed from science exposures. Moreover we present and implement a method to stack images of different quality resulting in an optimal S/N ratio for faint sky dominated {point sources}. The Js and K' band cover an area of about \\mbox{$200\\sq\\arcmin$} (1 patch) whereas the H band covers about $0.85\\sq\\degr$ (15 patches) in total. The 50\\% completeness limits are 22.67, $\\sim 21.9$, and 21.76 in the Js, H, and K' band, respectively. The number counts of all NIR bands nicely agree with the number counts taken from literature. Furthermore we present a deep and homogeneous i-band selected multi-waveband catalogue in the COSMOS field by combining publicly available u, B, V, r, i, z, and K bands with the H band. The clean catalogue with a formal 50\\% completeness limit for point sources of $i\\sim 26.7$ comprises about 290~000 galaxies with information in 8 passbands and covers an area of about $0.7\\sq\\degr$ (12 patches). We exclude all objects with corrupted magnitudes in only one of the filters from the catalogue in order to have a catalogue as homogeneous as possible. Photometric redshifts for all objects are derived and a comparison with 162 spectroscopic redshifts in the redshift range $ 0 \\lsim z \\lsim 3$ shows that the achieved accuracy of the photometric redshifts is \\mbox{$\\Delta z / (z_{spec}+1) \\lsim 0.035$} with only $\\sim 2$\\% outliers. Please note that in order to break the degeneracy between high redshift and low redshift solutions we included also the GALEX FUV and NUV filters in the photometric redshift estimation which considerably reduced the number of outliers. The multi-waveband catalogue including the photometric redshift information is made publicly available. The data can be downloaded from {\\verb http://www.mpe.mpg.de/~gabasch/COSMOS/ } We derive absolute UV magnitudes and a comparison in a magnitude-redshift diagram with the FDF shows good agreement. Moreover we investigate the evolution of the luminosity function evaluated in the restframe UV (1500~\\AA). We find a substantial brightening of M$^\\ast$ and a decrease of $\\phi^\\ast$ with redshift: from \\mbox{$\\langle z \\rangle\\sim 0.5$} to \\mbox{$\\langle z \\rangle\\sim 4.5$} the characteristic magnitude increases by about 3 magnitudes, whereas the characteristic density decreases by about 80 -- 90\\%. We compare the redshift evolution of the UV luminosity density in the COSMOS field and the FDF up to a redshift of $z\\sim 5$. Below a redshift of $z\\sim 2.5$ the mean UV luminosity density in COSMOS is systematically higher by about $1\\sigma$ if compared to the FDF. At $2.5 \\lsim z \\lsim 5$ both UV luminosity densities agree very well. It is worth noting the remarkably good agreement between the UV LF as well as the UV LD despite the fact, that the FDF is about 60 times smaller than the COSMOS field analysed here." }, "0710/0710.2464_arXiv.txt": { "abstract": "{} {We studied the temporal and spectral evolution of the synchrotron emission from the high energy peaked BL~Lac object \\1e.} {Two recent observations have been performed by the \\xmm\\,\\,and \\sw\\,satellites; we carried out X-ray spectral analysis for both of them, and photometry in optical-ultraviolet filters for the \\sw\\,\\,one. Combining the results thus obtained with archival data we built the long-term X-ray light curve, spanning a time interval of 26~years, and the Spectral Energy Distribution (SED) of this source.} {The light curve shows a large flux increasing, about a factor of six, in a time interval of a few years. After reaching its maximum in coincidence with the \\xmm\\,\\,pointing in December~2000 the flux decreased in later years, as revealed by \\sw\\,. The very good statistics available in the 0.5-10 keV \\xmm\\,\\,X-ray spectrum points out a highly significant deviation from a single power law. A log-parabolic model with a best fit curvature parameter of 0.25 and a peak energy at $\\sim$~1~keV describes well the spectral shape of the synchrotron emission. The simultaneous fit of \\sw\\,\\,UVOT and XRT data provides a milder curvature ($b\\sim0.1$) and a peak at higher energies ($\\sim15$~keV), suggesting a different state of source activity. In both cases UVOT data support the scenario of a single synchrotron emission component extending from the optical/UV to the X-ray band.} {New X-ray observations are important to monitor the temporal and spectral evolution of the source; new generation $\\gamma$-ray telescopes like AGILE and GLAST could for the first time detect its inverse Compton emission.} ", "introduction": "BL~Lac objects are thought to be radio-loud Active Galactic Nuclei (AGNs) observed in a direction very close to the axis of a relativistic jet outflowing from the inner nuclear region (\\cite{Urry95}). This interpretation could explain most of the characteristics of these sources like compact and flat-spectrum radio emission, superluminal motion revealed by VLBI imaging, high and variable radio and optical polarization, non-thermal continuum emission extending from radio to $\\gamma$-ray frequencies, an almost featureless optical spectrum and the fast variability at all frequencies. BL~Lac objects are generally characterized by a double bump structure in the broad band Spectral Energy Distribution (SED). The low frequency bump is attributed to synchrotron radiation emitted by relativistic electrons in the jet; inverse Compton scattering by the same electron population on the synchrotron radiation is thought to be at the origin of the high frequency bump. The peak of the first bump may vary in a rather wide range of frequencies: from the IR/optical band for the low energy peaked BL~Lacs (LBLs) to the UV/X-ray band for the high energy peaked BL~Lacs (HBLs) (\\cite{Giommi94}, \\cite{Padovani}). \\1e, also named BZB J1210+3929 in the recent Multifrequency Catalogue of Blazars (\\cite{BZcat}) is one of the X-ray selected BL~Lacertae objects of the \\ein\\,\\,Medium Sensitivity Survey (\\cite{EMSS}). It was discovered as a serendipitous source located about five arc-minutes north of one of the most intensively studied AGNs, the bright Seyfert galaxy NGC~4151. For this reason \\1e has been observed on many occasions by all the imaging X-ray instruments that have operated since the \\ein\\,\\,observatory. Despite its relatively high redshift (z=0.615) HST was able to detect the bright (M$_{\\rm R}$\\,=\\,$-$24.4) host galaxy which is of elliptical type (\\cite{Scarpa00}). In this paper we report the long-term X-ray light curve which spans over 26 years. We also report and compare the spectral analysis of the most recent observations carried out by the \\xmm\\,\\,(\\cite{jansen}) and \\sw\\,\\,(\\cite{gehrels}) satellites; we discuss the possibility to model the synchrotron emission of this source with a single log-parabolic model. The peak of the synchrotron component lies in the X-ray band, and for this reason the source can safely be classified as an HBL. We finally report the Spectral Energy Distribution (SED) of \\1e compiled from non-simultaneous multi-frequency archival data. ", "conclusions": "\\label{Discussion} Adding to multi-frequency archival data the results of our analysis we built the radio to X-ray SED of \\1e (Fig.\\ref{SED}) which follows the characteristic trend of extreme HBL sources (\\cite{Padovani}) with the synchrotron emission covering the entire frequency interval and peaking well into the X-ray band. The inverse Compton component is not detected and therefore is expected at higher energies. The \\sw\\,\\,satellite is sensitive both in the X-ray and in the optical-ultraviolet band. This provides the opportunity to test if the Spectral Energy Distribution of a source like \\1e can be explained by a single synchrotron component, or if multiple emission components are present. For this purpose we first fitted a log-parabolic model to simultaneous \\sw\\,\\,XRT and UVOT fluxes (solid line in Fig.~\\ref{combo}) and obtained a rather low value for the curvature parameter ($b\\sim0.1$) in contrast with the one obtained by fitting XRT data only, as reported in Table~\\ref{tab1}; the energy peak lies at about 15~keV or more. Comparing these results with those from \\xmm, \\sw\\,\\,evidently caught \\1e in a different state of activity, characterized by a milder curvature and a peak of the synchrotron component shifted to higher energies. Anyway firm conclusions cannot be drawn due to a relatively poor statistics and to the fact that the found $E_p$ value likely lies at energies higher than 10~keV, outside the range of \\sw\\,\\,XRT. Aware of the limits of using non simultaneous observations for a variable source, yet we veri\\-fied if \\sw\\,\\,UVOT and \\xmm\\,\\,MOS1 data are compatible with a log-parabolic model. We took points representing XSPEC model of observation~I and fit them on a wider frequency interval (dot-dashed line in Fig.\\,\\ref{combo}): UVOT points are sistematically above the extrapolation of the log-parabolic model. We took care to compare this result with the one obtained by fitting directly X-ray rebinned data and obtained an almost coincident result. Then we fitted both data sets with a log-parabola (dotted line in Fig.\\,\\ref{combo}) and estimated the curvature parameter: we obtained $b=0.18\\pm0.01$, a value at about 2$\\sigma$ of the one derived fitting only \\xmm\\,\\,data (see Table \\ref{tab1}). At this point we tested the log-parabolic model with $b$ fixed at 0.18 and found $\\chi^2_r/d.o.f. = 1.02/376$, a value practically coincident with the one obtained leaving $b$ free to vary, with only a small difference in the $a$ parameter. This result encouraged us in concluding that a single synchrotron component is in a good agreement with the emission observed. \\1e is certainly worth monitoring in the next years. New observations in the X-ray band would add essential information to the light curve behaviour: if any kind of regularity in flux increasing and fading should emerge, it would be possible to put physical constraints to establish the nature of the mechanism responsible for these variations. Moreover, observations with instruments which extend to higher energies than XRT, like those on board \\textit{Suzaku}, would extend the known Spectral Energy Distribution and would allow us to obtain a better parametrization of the spectral curvature. Finally, \\1e may be an interesting target for AGILE and the forthcoming GLAST $\\gamma$-ray mission which could detect for the first time the inverse Compton component of this HBL source." }, "0710/0710.0182_arXiv.txt": { "abstract": "Using high-resolution N-body simulations, we examine whether a major dry merger mitigates the difference in the radial density distributions between red and blue globular clusters (GCs). To this end, we study the relation between the density slope of the GCs in merger progenitors and that in a merger remnant, when the density distribution is described by $n_{\\rm GC}\\propto r^{-\\alpha}$. We also study how our results depend on the merger orbit and the size of the core radius of the initial GC density distribution. We find that a major dry merger makes the GC profile flatter, and the steeper initial GC profile leads to more significant flattening, especially if the initial slope is steeper than $\\alpha\\sim3.5$. Our result suggests that if there is a major dry merger of elliptical galaxies whose red GCs have a steeper radial profile than the blue GCs, as currently observed, and their slopes are steeper than $\\alpha\\sim3.5$, the difference in the slopes between two populations becomes smaller after dry mergers. Therefore, the observed slopes of red and blue GCs can be a diagnostic of the importance of dry merger. The current observational data show that the red and blue GCs have more comparable and shallower slopes in some luminous galaxies, which may indicate that they have experienced dry mergers. ", "introduction": "Globular clusters (GCs) in elliptical galaxies have been intensively studied in consideration of explaining the formation of both their host elliptical galaxies and GCs themselves \\citep[see][for a review]{araa06}. GCs are attractive as tracers of the star formation history of their host galaxies \\citep[e.g.][]{yi04,strader06}, because some properties of GC systems are correlated with the properties of their host galaxies \\citep[e.g.,][]{brodie91,djorgovski92}. It is thought that the formation of GCs is triggered by starburst accompanying gas-rich galaxy merging \\citep[e.g.][]{schweizer87,ashman92} or starburst that might happen with multiple dissipational collapses \\citep{forbes97}. Forming young star clusters are found and they are expected to become star clusters like current old GCs in local galaxy mergers \\citep{schweizer06}. An important aspect of GC systems in elliptical galaxies is a color bimodality \\citep[e.g.,][]{zepf93,geisler96,gebhardt99,larsen01,peng06}. It is also found that red GCs are more centrally concentrated than blue GCs \\citep[e.g.][]{forte05,bassino06b, tamura06}, which could put additional constraints on their formation scenario \\citep{bekki02}. The radial profile of each GC subpopulation is well described by a power-law distribution, especially at the outer radii, and the red GCs have a steeper slope than the blue GCs. Although the origin of this color bimodality is still uncertain, it is probably closely related to the formation history of their host elliptical galaxies \\citep[e.g.,][]{yoon06,strader07,kundu07}. Classically, three scenarios have been proposed to explain the color bimodality; major gas-rich mergers, in situ formation of multiple dissipational collapses, and dissipationless accretion. An explanation suggested by \\citet{ashman92} is that red GCs are metal-rich clusters which might be formed by gas-rich disk-disk mergers. Therefore, the red GCs might be younger than blue GCs. Meanwhile, \\citet{forbes97} explain that blue GCs might be formed in the first stage of dissipational collapse and red GCs might be formed after the truncation of the blue GC formation. Another explanation given by \\citet{cote98} includes accretion of blue GCs from small galaxies into the already formed red GCs. More recently, \\citet{beasley02} demonstrate that the bimodality can be explained in elliptical galaxy formation based on a hierarchical clustering scenario. On the other hand, recent research suggests that in the late stage of evolution, early-type galaxies might have experienced dry merging where merger progenitors do not have much gas, nor accompany star formation. The number density evolution of red galaxies has been discussed in observations of COMBO-17 \\citep{bell04} and DEEP2 surveys \\citep{faber05}, and such studies suggest that the density change can be understood by the dominance of dry merging after z $\\simlt$ 1 \\citep[but see also][]{yamada05,cimatti06,bundy07,scarlata07}. The evolution of galaxy clustering also implies late effects on the evolution of massive red galaxies from dry merging \\citep{white07}. Moreover, the observations show that dry merging does occur \\citep{vandokkum05,tran05,rines07}, while the observed features of galaxies are well explained in cosmological simulations \\citep[e.g.][]{kawata06}. Recent theoretical studies of dry merger simulations of ellipticals show that merger remnants maintain their properties on the fundamental plane and other scaling relations \\citep[e.g.][]{nipoti03,boylan05,robertson06,ciotti07}. The dry merging of binary ellipticals also can explain the formation of boxy-type ellipticals \\citep{naab06}. \\citet{bekki06a} demonstrate that the observed correlation between a spatial distribution of GCs and the total luminosity of ellipticals can be explained by sequential dissipationless major mergers, because the radial density profile of GCs progressively flatten after each major dry merger. This study raises the important question of whether or not the slopes of the density profiles of red and blue GCs persist after major dry merging. For example, we now consider that the density profiles of red and blue GCs in progenitor elliptical galaxies are described by $n_{\\rm GC, red}\\propto r^{-\\alpha_{\\rm p,red}}$ and $n_{\\rm GC, blue}\\propto r^{-\\alpha_{\\rm p,blue}}$, and these profiles in a major dry merger remnant become $n_{\\rm GC, red}\\propto r^{-\\alpha_{\\rm r,red}}$ and $n_{\\rm GC, blue}\\propto r^{-\\alpha_{\\rm r,blue}}$. The current observations suggest that $\\alpha_{\\rm r,red}>\\alpha_{\\rm r,blue}$. However, if a dry merger flattens the red GC density profile more than the blue GC density profile, $\\alpha_{\\rm p,red}-\\alpha_{\\rm r,red}>\\alpha_{\\rm p,blue}-\\alpha_{\\rm r,blue}$, in the remnant galaxy the difference between $\\alpha_{\\rm r,red}$ and $\\alpha_{\\rm r, blue}$ becomes smaller. In this case, the observed difference in the slopes of the density profiles of red and blue GCs in nearby ellipticals can be a valuable diagnostic for the importance of the dry merger in the evolution of ellipticals. To clarify this issue, the question becomes how the flattening of GC profiles during dry merging depends on the initial distributions of the GCs. We use numerical simulations of major dry mergers to study the dependence of GC distributions in merger remnants on the initial distributions in merger progenitors. Then, we can compare $\\alpha_{\\rm p,red}-\\alpha_{\\rm r,red}$ and $\\alpha_{\\rm p,blue}-\\alpha_{\\rm r,blue}$, for different sets of $\\alpha_{\\rm p,red}$ and $\\alpha_{\\rm p,blue}$. Since major dry mergers between two equal-mass merger progenitors must leave the most significant effects on merger remnants, compared with minor mergers, we study only equal-mass mergers in this paper. In \\S2, we explain details of our dry merger simulations and initial GC distributions. The changes in GC spatial distributions due to major mergers are shown in \\S3. We discuss the implication of our results in \\S4 that is followed by conclusion. ", "conclusions": "Our main result is that steeper initial GC profiles experience stronger flattening as shown in Figure \\ref{fig:alpha}. $\\alpha_{\\rm p} \\approx 3.5$ is in boundary between strong and weak flattening. The results imply that if the initial slopes of both red and blue GCs are steeper than $\\alpha_{\\rm p} \\approx 3.5$, the difference in the slopes between two populations of GCs will become much smaller, independent of merger orbits. In particular, even only one dry merger can make the slope flatter dramatically. Moreover, it also makes both slopes be around $\\alpha_{\\rm r} \\approx 3.5 - 4.5$. For example, if $\\alpha_{\\rm p,red}=5$ and $\\alpha_{\\rm p, blue}=4$, Figure \\ref{fig:alpha} suggests that $\\alpha_{\\rm r,red} \\approx 4.3$ and $\\alpha_{\\rm p, blue} \\approx 3.6$. The difference in the slope between red and blue GCs becomes significantly small. Therefore, the difference in the slopes of red and blue GCs can be a constraint of the number of major dry mergers. On the other hand, if the initial GC distribution has a slope shallower than $\\alpha_{\\rm p} \\approx 3.5$, the slope changes very little, and almost no change in the case of $\\alpha_{\\rm p} \\approx 2$. Therefore, if the initial distributions of red and blue GCs in merger progenitors are steeper than $\\alpha_{\\rm p} = 2$, and the galaxies experienced a number of major dry mergers, it leads the distributions of both red and blue GCs to become close to $\\alpha_{\\rm r} \\sim 2$ for both red and blue GCs, and the difference in the slope of spatial distributions becomes difficult to be measured. It is also worth noting that if the initial core size is larger for blue GCs than for red GCs, but their initial slopes are the same, a major dry merger can make the slope for the blue GCs shallower than for the red GCs, as shown in Figure \\ref{fig:r_c_dependence1}. Therefore, the final slope is not a simple function of the initial slope. Our results suggest that dry mergers make the slopes of the density profiles for red and blue GCs shallower and similar. Several studies \\citep[e.g.][]{kissler97,vandenbergh98,forbes05,lauer07, emsellem07} claim that the various observed properties for ellipticals show a transition around ${\\rm M_{V} \\sim -21}$, i.e. ${\\rm \\sim 10^{11}\\ M_{\\sun}}$. Some of these bimodalities have been interpreted in the picture of dry merger hypothesis for the growth of massive ellipticals \\citep[e.g.][]{capetti06}. For example, \\citet{emsellem07} show that slowly rotating ellipticals might be mainly affected by dry mergers. The slowly rotating ellipticals are more luminous than ${\\rm M_{B} \\approx -20.5}$, while less luminous ellipticals are fast rotators. It may indicate that less luminous galaxies have not experienced any major dry mergers. Therefore, it is interesting to examine the observed distributions of red and blue GCs as a function of the stellar mass of ellipticals. If more luminous galaxies have experienced dry mergers, our results predict that the slopes for red and blue GCs in bright galaxies are shallower and similar. Unfortunately, the current observational samples are not enough to test it statistically. Below we provide some discussion based on the current limited measurements. NGC 1399 is one of ellipticals whose spatial distributions of red and blue GCs are well-observed \\citep{dirsch03,bassino06b}. When assuming ${\\rm M/L_{V} = 5}$ and $B-V=0.9$ for an elliptical galaxy, as used in \\citet{forbes05}, the stellar mass of NGC 1399 is about ${\\rm 5 \\times\\ 10^{11}\\ M_{\\odot}}$ (${\\rm M_{B} = -21.8}$) \\citep{araa06}. The derived power-law indices of projected radial density profiles are $1.9\\pm0.06$ and $1.6\\pm0.10$ for red and blue GCs, respectively \\citep{bassino06b}. If the distributions are spherically symmetric, the three-dimensional radial distributions have power-law slopes of $\\alpha \\approx 2.9$ and $\\approx 2.6$ for red and blue GCs, respectively. Note that the real slope might be slightly steeper, because the distribution of GCs is not infinite. The slope difference between red and blue GC distributions is $\\Delta \\alpha \\approx 0.3$ while both GC populations show $\\alpha \\approx 3$. In addition to NGC 1399, NGC 1407 is a brightest group galaxy with ${\\rm M_V=-21.86}$, and has a bimodal color distribution of GCs. \\citet{forbes06} report that the projected slopes of the red and blue GC density profiles are respectively $1.50\\pm0.05$ and $1.65\\pm0.29$, i.e., expected three dimensional slopes of $\\alpha_{\\rm red}=2.50$ $\\alpha_{\\rm blue}=2.65$, which are statistically the same as each other. On the other hand, NGC 1374 and NGC 1379 are less luminous than NGC 1399, having ${\\rm M_{V} = -20.4}$ and $-20.6$, respectively. The red and blue GCs of the two galaxies are studied in \\citet{bassino06a}. The derived projected density distributions of red and blue GCs in NGC 1374 have slopes of 3.2 and 2.3. Hence, the expected slopes in their three-dimensional density profiles are $\\alpha_{\\rm red}\\sim 4.2$ and $\\alpha_{\\rm blue}\\sim3.3$. The $\\Delta \\alpha$ in NGC 1379 is $\\sim 0.6$, and the red and blue GCs have projected power-law slopes of $\\sim 2.9$ and $\\sim 2.3$, respectively. The low-luminosity galaxies have a larger difference in the density slopes of the red and blue GCs, and their slopes are steeper than the more luminous galaxies discussed above. Combining with our results, these four sample suggests that luminous galaxies, like NGC 1399 and NGC 1407, may have experienced some dry mergers, while less luminous galaxies, such as NGC 1374 and NGC 1379, may have not been formed through dry mergers, which is consistent with a scenario that the significance of dry mergers depends on the stellar mass. However, we also find some giant galaxy that does not follow the above trend. Recently, \\citet{tamura06} measured the density profiles of red and blue GCs for giant ellipticals: M87 and NGC 4552. We have fitted the projected density profile in Figure 5 of their paper by a power-law profile, and the derived power-law index is 2.4 and 1.4 for red and blue GCs in M87 and 1.8 and 1.2 for red and blue GCs in NGC 4552, respectively. Although these giant ellipticals have relatively shallower slopes, the difference in slopes of red and blue GCs is significant. According to our result, these giant ellipticals are unlikely to have experienced significant dry mergers. Some giants might form without dry mergers, or some other factor, such as a number of minor mergers, might affect the slopes of GCs. Again, so far there are not many studies which derive the density distributions of both red and blue GCs. This problem limits the application of our results to the current observational data. However, we expect that future observations will improve the statistics for nearby ellipticals, and our results will add an important framework to interpret the galaxies' evolutionary history. It will be also valuable to compare spatial distributions of blue and red GCs with other expected properties from dry mergers such as surface brightness profile." }, "0710/0710.5857_arXiv.txt": { "abstract": "The classical picture of GUT baryogenesis has been strongly modified by theoretical progress concerning two nonperturbative features of the standard model: the phase diagram of the electroweak theory, and baryon and lepton number changing sphaleron processes in the high-temperature symmetric phase of the standard model. We briefly review three viable models, electroweak baryogenesis, the Affleck-Dine mechanism and leptogenesis and discuss the prospects to falsify them. All models are closely tied to the nature of dark matter, especially in supersymmetric theories. In the near future results from LHC and gamma-ray astronomy will shed new light on the origin of the matter-antimatter asymmetry of the universe. ", "introduction": "The cosmological matter-antimatter asymmetry can be dynamically generated if the particle interactions and the cosmological evolution satisfy Sakharov's conditions \\cite{sak67}, \\begin{itemize} \\item baryon number violation, \\item $C$ and $C\\!P$ violation, \\item deviation from thermal equilibrium. \\end{itemize} Although the baryon asymmetry is just a single number, it provides an important connection between particle physics and cosmology. In his seminal paper, 40 years ago, Sakharov not only stated the necessary conditions for baryogenesis, he also proposed a specific model. The origin of the baryon asymmetry were $C\\!P$ violating decays of superheavy `maximons' with mass $\\mathcal{O}(M_\\mathrm{P})$ at an initial temperature $T_i \\sim M_\\mathrm{P}$. The $C\\!P$ violation in maximon decays was related to the $C\\!P$ violation observed in $K^0$-decays, and the violation of baryon number led to a proton lifetime $\\tau_p > 10^{50}\\ \\mathrm{years}$, much larger than current estimates in grand unified theories. \\begin{figure}[t] \\includegraphics[height=6cm]{fig3v}\\hspace{1cm} \\includegraphics[height=6cm,width=8cm]{pot} \\caption{{\\it Left:} Critical temperature $T_c$ of the electroweak transition as function of $R_{HW}=m_H/m_W$; from \\cite{cfh98}. {\\it Right:} Effective potential of the Higgs field $\\varphi$ at temperature $T>T_c$. \\label{fig:ew}} \\end{figure} At present there exist a number of viable scenarios for baryogenesis. They can be classified according to the different ways in which Sakharov's conditions are realized. In grand unified theories baryon number ($B$) and lepton number ($L$) are broken by the interactions of gauge bosons and leptoquarks. This is the basis of classical GUT baryogenesis (cf.~\\cite{kt90}). In a similar way, lepton number violating decays of heavy Majorana neutrinos lead to leptogenesis \\cite{fy86}. In the simplest version of leptogenesis the initial abundance of the heavy neutrinos is generated by thermal processes. Alternatively, heavy neutrinos may be produced in inflaton decays or in the reheating process after inflation. Because in the standard model baryon number, $C$ and $C\\!P$ are not conserved, in principle the cosmological baryon asymmetry can also be generated at the electroweak phase transition \\cite{krs85}. A further mechanism of baryogenesis can work in supersymmetric theories where the scalar potential has approximately flat directions. Coherent oscillations of scalar fields can then generate large asymmetries \\cite{ad85}. The theory of baryogenesis crucially depends on nonperturbative properties of the standard model, first of all the nature of the electroweak transition. A first-order phase transition yields a departure from thermal equilibrium. Fig.~1 shows the phase diagram of the electroweak theory, i.e. the critical temperature in units of the Higgs mass, $T_c/m_H$, as function of the Higgs mass in units of the W-boson mass, $R_{HW}=m_H/m_W$ \\cite{cfh98,lr98}. For small Higgs masses the phase transition is first-order; above a critical Higgs mass, $m_H > m_H^c \\simeq 72$~GeV, it turns into a smooth crossover \\cite{bp94,klx96}. This upper bound for a first-order transition has to be compared with the lower bound from LEP, $m_H > 114$~GeV. Hence, there is no departure from thermal equilibrium at the electroweak transition in the standard model. The second crucial nonperturbative aspect of baryogenesis is the connection between baryon number and lepton number in the high-temperature, symmetric phase of the standard model. Due to the chiral nature of the weak interactions $B$ and $L$ are not conserved \\cite{tho76}. At zero temperature this has no observable effect due to the smallness of the weak coupling. However, as the temperature reaches the critical temperature $T_c$ of the electroweak phase transition, $B$ and $L$ violating processes come into thermal equilibrium \\cite{krs85}. The rate of these processes is related to the free energy of sphaleron-type field configurations which carry topological charge. In the standard model they lead to an effective interaction of all left-handed fermions \\cite{tho76} (cf.~Fig.~2), \\begin{equation} O_{B+L} = \\prod_i \\left(q_{Li} q_{Li} q_{Li} l_{Li}\\right)\\; , \\end{equation} which violates baryon and lepton number by three units, \\begin{equation} \\Delta B = \\Delta L = 3\\;. \\label{sphal1} \\end{equation} The sphaleron transition rate in the symmetric high-temperature phase has been evaluated by combining an analytical resummation with numerical lattice techniques \\cite{bmr00}. The result is, in accord with previous estimates, that $B$ and $L$ violating processes are in thermal equilibrium for temperatures in the range \\begin{equation} T_{EW} \\sim 100\\ \\mbox{GeV} < T < T_{SPH} \\sim 10^{12}\\ \\mbox{GeV}\\;. \\end{equation} \\begin{figure}[t] \\includegraphics[height=6cm]{lg1} \\caption{One of the 12-fermion processes which are in thermal equilibrium in the high-temperature phase of the standard model. \\label{fig:sphal} } \\end{figure} Sphaleron processes have a profound effect on the generation of the cosmological baryon asymmetry. An analysis of the chemical potentials of all particle species in the high-temperature phase yields the following relation between the baryon asymmetry and the corresponding $L$ and $B-L$ asymmetries, \\begin{equation}\\label{basic} \\langle B\\rangle_T = c_S \\langle B-L\\rangle_T = {c_S\\over c_S-1} \\langle L\\rangle_T\\;. \\end{equation} Here $c_S$ is a number ${\\cal O}(1)$. In the standard model with three generations and one Higgs doublet one has $c_s= 28/79$. We conclude that lepton number violation is necessary in order to generate a cosmological baryon asymmetry\\footnote{In the case of Dirac neutrinos, which have extremely small Yukawa couplings, one can construct leptogenesis models where an asymmetry of lepton doublets is accompanied by an asymmetry of right-handed neutrinos such that the total lepton number is conserved and $\\langle B-L \\rangle_T = 0$ \\cite{dlx00}.}. However, it can only be weak, because otherwise any baryon asymmetry would be washed out. The interplay of these conflicting conditions leads to important contraints on neutrino properties and on possible extensions of the standard model in general. \\begin{figure}[t] \\includegraphics[height=6cm]{ewbg} \\caption{Sketch of nonlocal electroweak baryogenesis. From \\cite{ber02}. } \\end{figure} ", "conclusions": "40 years after Sakharov's work on the cosmological matter-antimatter asymmetry we have several viable models of baryogenesis, the most predictive ones being electroweak baryogenesis and leptogenesis. In fact, based on our theoretical understanding of the electroweak phase diagram, electroweak baryogenesis in the standard model has already been excluded by the LEP bound on the Higgs mass. Supersymmetric electroweak baryogenesis will soon be tested at the LHC. Detailed studies of the nonequilibrium leptogenesis process have led to the preferred neutrino mass window $10^{-3}\\ {\\rm eV} < m_i < 0.1\\ {\\rm eV}$ in the simplest scenario with hierarchical heavy neutrinos. The consistency with the experimental evidence for neutrino masses has dramatically increased the popularity of the leptogenesis mechanism. It is exciting that new experiments and cosmological observations will probe the absolute neutrino mass scale in the coming years. However, more work is needed on the full quantum mechanical treatment of leptogenesis, in particular the flavour dependence. All baryogenesis mechanisms are closely related to the nature of dark matter. A discovery of the standard supergravity scenario at LHC could be consistent with electroweak baryogenesis but would rule out the simplest version of thermal leptogenesis. On the other hand, evidence for gravitino dark matter can be consistent with leptogenesis. Finally, the discovery of macroscopic dark matter like Q-balls would point towards nonperturbative dynamics of scalar fields in the early universe and therefore favour Affleck-Dine baryogenesis." }, "0710/0710.2193_arXiv.txt": { "abstract": "Rapid mass transfer in a binary system can drive the accreting star out of thermal equilibrium, causing it to expand. This can lead to a contact system, strong mass loss from the system and possibly merging of the two stars. In low metallicity stars the timescale for heat transport is shorter due to the lower opacity. The accreting star can therefore restore thermal equilibrium more quickly and possibly avoid contact. We investigate the effect of accretion onto main sequence stars with radiative envelopes with different metallicities. We find that a low metallicity ($Z<10^{-3}$) $4\\Msun$ star can endure a 10 to 30 times higher accretion rate before it reaches a certain radius than a star at solar metallicity. This could imply that up to two times fewer systems come into contact during rapid mass transfer when we compare low metallicity. This factor is uncertain due to the unknown distribution of binary parameters and the dependence of the mass transfer timescale on metallicity. In a forthcoming paper we will present analytic fits to models of accreting stars at various metallicities intended for the use in population synthesis models. ", "introduction": "The majority of stars are found in binaries, many of which can interact, for example by exchanging mass, resulting in an evolution very distinct from isolated stars. Although the fraction of stars in binaries might be different for earlier generations of stars, formed in metal-poor environments, they are certainly worth a systematic study. In this work we discuss the effect of accretion onto main sequence stars as function of metallicity. In a second contribution to these proceedings we discuss how the ranges for different cases of mass transfer depend on metallicity \\citep{caseABC07}. Mass transfer takes place when one of the stars exceeds a certain critical radius, the Roche lobe radius. The first phase of mass transfer is usually so fast that the accreting star is driven out of thermal equilibrium and expands \\citep{Benson70,Yungelson73}, potentially so much that the two stars come into contact. Contact binaries are not well understood, but they probably involve strong mass loss from the system and merging of the two stars. Instead of using full binary evolution models, we choose to reduce the large parameter space of binaries by studying the behavior of models of isolated stars under controlled conditions. We follow the approach of \\citet{Kippenhahn+Meyer77} and \\citet{Neo+ea77}, who studied the evolution of the radius of accreting main sequence stars. We extend their work, to study the effect of metallicity, using up-to-date input physics. Here we present the results of an exploratory study and its possible implications for binaries at low metallicity. In a forth coming paper we will present analytic fits to a finer grid of models, intended for the use in population synthesis models. Expansion due to thermal timescale mass transfer is commonly neglected or taken into account using a simple approximate criterion. Our fits will provide an improvement, which is easy to implement. \\begin{figure} \\includegraphics[ angle=-90, width=\\columnwidth]{deminkea_poster1_fig1.ps} \\caption{Radius versus mass of an initially 4 \\Msun star at solar metallicity accreting with different accretion rates. \\label{R_vs_m} } \\end{figure} ", "conclusions": "We find that a $4\\Msun$ low metallicity star ($Z<10^{-3}$) can endure a 10 to 30 times higher accretion rate before it expands to a certain radius compared to solar metallicity stars. This suggests that at low metallicity fewer binaries come into contact during rapid mass transfer. A rough estimate based on very simplified assumptions indicate that the effect could be up to a factor 2, i.e. only half as many may binaries evolve into contact at low metallicity(Z < 10{-3}) compared to solar metallicity. This factor is uncertain and probably only an upper limit as it depends on the initial distribution of binary parameters, on how the typical mass transfer rate depends on metallicity \\citep[it is probably higher at low metallicity, e.g.][]{Langer_ea00} and the binary parameters, on the specific entropy of the accreted material and on the efficiency of mass transfer \\citep[see for example][]{DeMink+ea07}. In a forthcoming paper we will present analytic fits to our models intended for the use in population synthesis models. \\begin{theacknowledgments} We would like to thank Peter Eggleton for providing his stellar evolution code, John Eldridge for the opacity tables and Rob Izzard for interesting discussions and suggestions. \\end{theacknowledgments}" }, "0710/0710.3706_arXiv.txt": { "abstract": " ", "introduction": "The Magellanic Clouds (MCs) are interacting SBm galaxies similar to many that exist in Universe. They are the largest neighbouring satellites of the Milky Way, reflecting a typical environment of a large galaxy surrounded by satellites. They contain stars which are as old as the Universe as well as newly forming and this extended range of star formation is a highly valuable source to understand the process of formation and evolution of galaxies in general. The MCs are overall more metal poor than the Galaxy and therefore may hold information about the Universe at its early stages. They are located at a fairly well known distance, which makes it easier to measure details of their stellar component and structure. They are also fortunately located in a region of sky only lightly affected by Galactic reddening, which translates into the capability of detecting their faint stellar populations. The MCs belong to a complex system, the Magellanic System, which has in total four distinct components: the Large Magellanic Cloud (LMC), the Small Magellanic Cloud (SMC), the Bridge connecting the two Clouds and the Stream attached to the SMC. The latter two are predominantly formed of gas and are of tidal origin. ", "conclusions": "The Magellanic System has yet many challenging aspects that new surveys, with the increased quality of the coming data and new theoretical models and their ability to explain detail observations, aim to resolve in the next decade. Prior to new facilities like GAIA, JWST and ALMA we need to exploit data from VISTA and similarly powerful telescopes at other wavelengths. Surveys like VMC will provide unique and high quality data for science and training of young astronomers." }, "0710/0710.3530_arXiv.txt": { "abstract": "We present the first measurements of the angular correlation function of galaxies selected in the far (\\fuvcenter) and near (\\nuvcenter) Ultraviolet from the \\galex survey fields overlapping SDSS DR5 in low galactic extinction regions. The area used covers $120$ sqdeg (\\galex - MIS) down to magnitude AB $=22$, yielding a total of 100,000 galaxies. The mean correlation length is $\\sim3.7 \\pm 0.6$~Mpc and no significant trend is seen for this value as a function of the limiting apparent magnitude or between the \\galex bands. This estimate is close to that found from samples of blue galaxies in the local universe selected in the visible, and similar to that derived at $z\\simeq3$ for LBGs with similar rest frame selection criteria. This result supports models that predict anti-biasing of star forming galaxies at low redshift, and brings an additional clue to the downsizing of star formation at $z<1$. ", "introduction": "In the current paradigm of structure formation, the bulk of the most massive systems form in a cold dark matter-dominated universe by the merging of less massive units formed earlier. In parallel to this hierarchical evolution, recent observations point to the so-called ``downsizing'', namely the fact that in galaxies having high baryonic masses the bulk of stars formed at high redshift ($z \\gtrsim 1$), while in galaxies having low baryonic masses the bulk of stars formed at lower redshift \\citep[][and also \\citet{DeLucia_2006} and \\citet{Neistein_2006} for results from simulations]{Cowie_1996, Heavens_2004, Bundy_2006, Jimenez_2005}. The star formation efficiency shows a strong decline at $0$11--28\\,AU, well outside the location of Jupiter and Saturn in our Solar System. Even in other systems these outer radii are found to be devoid of giant planets (e.g. \\citealt{kasper07}). This fact suggests that the formation of terrestrial and giant planets may proceed undisturbed in disks around medium--separation binaries even if these disks are constrained in size. %Although disks in binaries are constrained in size, the formation of terrestrial and giant planets could then proceed undisturbed if the circumstellar material is similar in surface density and temperature to that in disks around single stars. These parameters are difficult to constrain from observations, but there are indications of a similar pathway for planet formation in disks around single and medium--separation binaries. Early investigations of young TTSs found no significant difference in the frequency of near-- and mid--infrared excess emission between single and binary star systems (e.g. \\citealt{simonprato95,jensen96}). With the 60\\micron{} IRAS flux probing dust $\\la$10\\,AU from the central star, these measurements demonstrate that binary systems as often have disks as single stars do. Recently \\citet{monin07} analyzed the separation distributions of binaries with and without disks and found no statistical difference. Since most of their binaries have projected separations $>$20\\,AU, their result shows that medium--and wide-- separation binaries do not have a significant effect on the circumstellar disk lifetime. Our work indicates that these disks also evolve in a similar way. The extent of dust processing in the disk surface layer and the degree of dust settling in binary disk systems do not statistically differ from those in disks around single stars. This suggests that the first few Myr of disk evolution in the terrestrial (and maybe out to the giant) planet--forming region are not affected by medium--separation stellar companions. Whether the disk evolution proceeds undisturbed for tens of millions of years until planets are fully formed cannot yet be assessed observationally. \\citet{bouwman06} estimate a mean disk dispersion timescale of $\\sim$5\\,Myr for close ($\\le$4AU) binaries in contrast to a timescale of $\\approx$9\\,Myr for single star systems. They argue that the time available to form planets in close binary systems is considerably shorter than that in disks around single stars, which may inhibit planet formation. The only two medium--separation binaries in their sample hint for a disk dispersal timescale comparable to that of single stars suggesting a similar disk evolution for single and medium--separation binary systems over the first $\\sim$\\,10\\,Myr. Exoplanet surveys offer us a glimpse into the frequency and properties of giant planets in multiple star systems. Recently \\citet{eu07} reported 42 planets orbiting binary and multiple stars (see, their Table 1). \\citet{bondes07} analyze a subsample of radial velocity planet host stars with uniform planet detectability and demonstrate that the overall frequency of giant planets in binaries is not statistically different from that of planets in single stars. However, they find indications for a lower frequency of radial velocity planets in the subgroup of close-- and medium--separation binaries ($<\\,50-100$\\,AU). In a complementary study, \\citet{desbar07} find that the mass distribution of planets in binaries with separations $<\\,300-500$\\,AU is statistically different from that around wider binaries and single stars: Massive planets in short--period orbits are found predominantly around close-- and medium--separation binaries. Taken together, the results from the frequency and properties of exoplanets suggest that a stellar companion with separation less than a few hundred AU affects giant planet formation and/or the subsequent migration. Numerical simulations seem to support this notion. \\citet{kley00} shows that a fairly eccentric ($e_{\\rm bin}=0.5$) stellar companion at 50--100\\,AU enhances the growth rate of a Jupiter mass planet embedded in a circumstellar disk and makes its inward migration more rapid. Recently, \\citet{kn07} confirm these trends by following the evolution of a 30\\,M$_\\earth$ protoplanet in a disk truncated by a stellar companion at 18.5\\,AU and $e_{\\rm bin}=0.36$, like the $\\gamma$~Cep binary system. Our study shows that the early evolution of protoplanetary disks surrounding binary stars is similar to that in single stars indicating that that the differences in the exoplanet properties arise in the later stages of their formation and/or migration. Whether terrestrial planet formation is also affected by medium--separation binaries cannot be yet addressed observationally. Our study shows that the initial dust processing is not impacted by the presence of a stellar companion. Based on the fact that the build--up of planetesimals as large as the $\\sim$500--km Vesta has occurred in the first 3.8$\\pm$1.3~Myr of the Solar nebula \\citep{kleine02}, it is reasonable to speculate on the basis of our study that the formation of planetesimals in binary and single systems proceed along, if not on identical avenues. Another indication supporting this suggestion comes from the finding of a similar incidence of debris disks in Gyr--old single and binary stars \\citep{trilling06}. If the debris dust is produced by colliding asteroids, then the similar rate of debris dust in binaries implies that planetesimal formation is not inhibited by the presence of stellar companions. Recent simulations of the later stages of terrestrial planet formation show that rocky planets can form in a wide variety of binary systems \\citep{quinta07}. The binary periastron is the most important parameter in limiting the number of forming planets and their range of orbits. \\citet{quinta07} show that binaries with periastron $\\ga$10\\,AU, comprising most of the medium--separation binaries investigated in this paper, can form terrestrial planets over the entire range of orbits allowed for single stars. As a result more than 50\\% of the binary systems in the Milky Way \\citep{dm91} are wide enough to allow the formation of Earth--like planets. \\subsection{The diversity in silicate features and SEDs} Although small sample statistics suggested a correlation between stellar multiplicity and initial dust processing \\citep{meeus03,sterzik04,sic07}, our study demonstrates that medium--separation stellar companions do not appreciably affect the growth and crystallization of dust in circumstellar disks. Given the criteria applied to select our samples, we can also exclude that age, spectral type, and stellar environment can account for the large variety of observed silicate emission features and SED slopes in our study. There may be several other factors contributing to this diversity that will be fully explored in an upcoming contribution. In the following we briefly mention two of them: Turbulence in circumstellar disks not only drives the accretion of gas onto the central star but also replenishes the disk atmosphere with more grains that can be larger in size. If the grains inferred from the 10\\,\\micron{} silicate emission feature reflect the level of disk turbulence, the strength of the features should depend on the stellar accretion rates. \\citet{sic07} note that stars with strong features tend to have large accretion rates in their sample of several Myr old intermediate-- and low--mass stars. This trend may be the result of turbulence determining the grain population in the disk atmosphere. Alternatively, the trend could be due to the more massive stars (that have typically larger accretion rates) in their sample heating larger disk area and thus producing stronger silicate emission features (see, e.g. \\citealt{kessler07}). The tentative correlation seen in the sample of \\citet{sic07} needs to be confirmed using a larger and more homogeneous sample of stars with well--determined accretion rates. %Stellar high--energy photons (UV and X--rays) may also affect the distribution of grain sizes and the composition of solids in circumstellar disks. Low--mass TTs are strong emitters of relative hard KeV X-rays (Wolk et al. 2005) that could evaporate the grain mantle or the complete grain. Voit (1992) calculated that silicate ?? grains with a radius $<$10\\AA evpaorates completed when subject to an X-ray energu of zz, which could be reached in the disk atmospheres for grains closer than yy from the central star. This could produce a grain size distribution biased toward larger grains. I DON'T UDERSTAND THE DRAINE A EQUANTION IT GIVES UNREASONABLE RESULTS!!! It says that 50um gains can survive (not evaporate as close as 2 stellar radii.) %Low-mass TTSs also show for about 25\\% of their time intense X--ray flares that can carry energies as high as a few MeV (Wolk et al. 2005). Such flares could melt and crystallize grains and may be responsible for the flash melted chondrules in meteorites (e.g. Shu et al. 2001). To asses the effect of stellar high--energy photons on small grains it would be interesting to correlate stellar UV/X-ray variability with changes in the strength and shape of the 10\\,\\micron{} silicate emission features. Different initial conditions for the collapsing cores may also leave their imprints on the formation and evolution of circumstellar disks. This possibility has been explored by \\citet{dullemond06} to explain crystallization of dust grains in the early stages of disk evolution. In their model the level of crystallinity depends crucially on the rotation rate of the collapsing cloud core because this determines the radius at which the infalling matter reaches the disk: rapidly rotating clouds would evolve into disks with low crystallinity, while slowly rotating clouds into disks with high crystallinity. In this paper we explored the effect of a stellar companion on the initial growth and settling of dust grains in circumstellar disks. We constructed two large samples of disks around single and binary TTSs with a narrow age spread and a spectral type distribution for the single stars identical to that of the primary stars in the binary sample. We used the strength of the 10\\,\\micron{} silicate emission feature derived from IRS/{\\it Spitzer} spectra as a proxy for grain growth and the SED slope of circumstellar disks as a proxy for dust settling. Our results can be summarized as follows: \\\\ \\smallskip -- there is no statistically significant difference between the distribution of 10\\,\\micron{} silicate emission features from single and binary systems. \\\\ -- the distribution of disk flaring is indistinguishable between the single and binary system samples. \\\\ \\smallskip These results show that stellar companions at projected separations of $\\ga$\\,10\\,AU do not appreciably affect the degree of crystallinity nor the degree of grain growth. Based on the combination of these and other results we argue that the formation of planetesimals and possibly terrestrial planets is not inhibited in a circumstellar disk perturbed by a medium--separation stellar companion. %%%%%%%%%% Single star spectra \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f8.ps} \\caption{Infrared spectra for the sample of single stars.\\label{s1}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f9.ps} \\caption{Infrared spectra for the sample of single stars.\\label{s2}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f10.ps} \\caption{Infrared spectra for the sample of single stars.\\label{s3}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f11.ps} \\caption{Infrared spectra for the sample of single stars.\\label{s4}} \\end{figure*} %%%%%%%%%% Binary star spectra \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f12.ps} \\caption{Infrared spectra for the sample of binary stars.\\label{b1}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f13.ps} \\caption{Infrared spectra for the sample of binary stars.\\label{b2}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f14.ps} \\caption{Infrared spectra for the sample of binary stars.\\label{b3}} \\end{figure*} \\begin{figure*} \\includegraphics[angle=90,scale=0.7]{f15.ps} \\caption{Infrared spectra for the sample of binary stars.\\label{b4}} \\end{figure*}" }, "0710/0710.1473_arXiv.txt": { "abstract": "In this contribution we study integrated properties of dynamically segregated star clusters. The observed core radii of segregated clusters can be 50\\% smaller than the ``true'' core radius. In addition, the measured radius in the red filters is smaller than those measured in blue filters. However, these difference are small ($\\lesssim10\\%$), making it observationally challenging to detect mass segregation in extra-galactic clusters based on such a comparison. Our results follow naturally from the fact that in nearly all filters most of the light comes from the most massive stars. Therefore, the observed surface brightness profile is dominated by stars of similar mass, which are centrally concentrated and have a similar spatial distribution. ", "introduction": " ", "conclusions": "" }, "0710/0710.0063_arXiv.txt": { "abstract": "An investigation of the dynamo instability close to the threshold produced by an ABC forced flow is presented. We focus on the on-off intermittency behavior of the dynamo and the counter-effect of the Lorentz force in the non-linear stage of the dynamo. The Lorentz force drastically alters the statistics of the turbulent fluctuations of the flow and reduces their amplitude. As a result much longer burst (on-phases) are observed than what is expected based on the amplitude of the fluctuations in the kinematic regime of the dynamo. For large Reynolds numbers, the duration time of the ``On'' phase follows a power law distribution, while for smaller Reynolds numbers the Lorentz force completely kills the noise and the system transits from a chaotic state into a ``laminar'' time periodic flow. The behavior of the On-Off intermittency as the Reynolds number is increased is also examined. The connections with dynamo experiments and theoretical modeling are discussed. \\pacs{47.65.-d,47.20.Ky,47.27.Sd,52.65.Kj} ", "introduction": "Dynamo action, the self amplification of magnetic field due to the stretching of magnetic field lines by a flow, is considered to be the main mechanism for the generation of magnetic fields in the universe \\cite{mhdbooks}. To that respect many experimental groups have successfully attempted to reproduce dynamos in liquid sodium laboratory experiments \\cite{gailitis2000,gailitis2001,gailitis2004,muller2000,stieglitz2001,monchaux2007,berhanu2007}. The induction experiments \\cite{odier1998,peffley2000a,peffley2000b,frick2002,bourgoin2002,nornberg2006a,nornberg2006b,stepanov2006,volk2006,bourgoin2006} studying the response of an applied magnetic field inside a turbulent metal liquid represent also a challenging science. With or without dynamo instability the flow of a conducting fluid forms complex system, with a large degree of freedoms and a wide branch of non linear behaviors. In this work we focus on one special behavior: the On-Off intermittency or blowout bifurcation \\cite{pomeau1980,platt1993}. On-off intermittency is present in chaotic dynamical systems for which there is an unstable invariant manifold in the phase space such that the unstable solutions have a growth rate that varies strongly in time taking both positive and negative values. If the averaged growth rate is sufficiently smaller than the fluctuations of the instantaneous growth rate, then the solution can exhibit on-off intermittency where bursts of the amplitude of the distance from the invariant manifold are observed (when the growth rate is positive) followed by a decrease of the amplitude (when the growth rate is negative). (See \\cite{sweet2001a,sweet2001b} for a more precise definition). On-Off intermittency has been observed in different physical experiments including electronic devices, electrohydrodynamic convection in nematics, gas discharge plasmas, and spin-wave instabilities \\cite{phys-onoff}. In the MHD context, near the dynamo instability onset, the On-Off intermittency has been investigated by modeling of the Bullard dynamo \\cite{leprovost2006}. Using direct numerical simulation \\cite{sweet2001a,sweet2001b} were able to observe On-Off intermittency solving the full MHD equations for the ABC dynamo, (here we present an extended work of this particular case). On-Off intermittency has also been found recently for a Taylor-Green flow \\cite{ponty2007b}. Finally, recent liquid metal experimental results (VKS) \\cite{pinton2007} show some intermittent behavior, with features reminiscent of on-off self-generation that motivated our study. For the MHD system we are investigating the evolution of the magnetic energy $E_b=\\frac{1}{2}\\int {\\bf b}^2 dx^3$ is given by $\\partial_t { E_b} = \\int {\\bf b( \\cdot b \\nabla) u - \\eta (\\nabla b)^2} dx^3$. If the velocity field has a chaotic behavior in time the right hand side of the equation above can take positive or negative values and can be modeled as multiplicative noise. A simple and proved very useful way to model the behavior of the magnetic field during the on-off intermittency is using a stochastic differential equation (SDE-model) \\cite{pomeau1980,platt1993,fujisaka,Yu1990,ott1004,Platt1994,Heagy1994,Venka1995,Venka1996,aumaitre2006,aumaitre2005}: \\begin{equation} \\partial_t E_b = (a+\\xi) E_b - NL(E_b) \\label{SDE} \\end{equation} where $ E_b$ is the magnetic energy, $a$ is the long time averaged growth rate, $\\xi$ models the noise term typically assumed to be white (see however \\cite{aumaitre2006,aumaitre2005}) and of amplitude $D$ such that $\\langle \\xi(t)\\xi(t') \\rangle = 2D\\delta(t-t')$. $NL$ is a non-linear term that guaranties the saturation of the magnetic energy to finite values typically taken to be $NL(X)=X^3$ for investigations of supercritical bifurcations or $NL(X)=X^5-X^3$ for investigations of subcritical bifurcations. Alternative, an upper no-flux boundary is imposed at $E_b=1$. In all these cases (independent of the non-linear saturation mechanism) the above SDE leads to the stationery distribution function that for $020$ On-Off intermittency was observed but with long durations of the ``on\" phases that have a power law distribution. These long ``on\" phases result in a pdf that peaks at finite values of $E_b$. This peak can be attributed to the presence of a subcritical instability or to the quenching of the hydrodynamic ``noise\" at the nonlinear stage or possibly a combination of the two. In principle the SDE model (eq.\\ref{SDE}) can be modified to include these two effects: a non-linear term that allows for a subcritical bifurcation and a $E_b$ dependent amplitude of the noise. There many possibilities to model the quenching of the noise, however the nonlinear behavior might not have a universal behavior and we do not attempt to suggest a specific model. The relative range of the On-Off intermittency was found to decrease as the Reynolds number was increased possibly reaching an asymptotic regime. However the limited number of Reynolds numbers examined did not allow us to have a definite prediction for this asymptotic regime. This question is of particular interest to the dynamo experiments \\cite{gailitis2000,gailitis2001,gailitis2004,muller2000,stieglitz2001,monchaux2007,berhanu2007} that until very recently \\cite{pinton2007} have not detected On-Off intermittency . There are many reasons that could explain the absence of detectable On-Off intermittency in the experimental setups, like the strong constrains imposed on the flow \\cite{gailitis2004,muller2000} that do not allow the development of large scale fluctuations or the Earths magnetic field that imposes a lower threshold for the amplitude of the magnetic energy. Numerical investigations at higher resolution and a larger variety of flows or forcing would be useful at this point to obtain a better understanding." }, "0710/0710.0580_arXiv.txt": { "abstract": "{Using lattice techniques we investigate the generation of long range cosmological magnetic fields during a cold electroweak transition. We will show how magnetic fields arise, during bubble collisions, in the form of magnetic strings. We conjecture that these magnetic strings originate from the alignment of magnetic dipoles associated with EW sphaleron-like configurations. We also discuss the early thermalisation of photons and the turbulent behaviour of the scalar fields after tachyonic preheating.} \\FullConference{The XXV International Symposium on Lattice Field Theory\\\\ July 30 - August 4 2007\\\\ Regensburg, Germany} \\begin{document} ", "introduction": "There have been many theoretical attempts to explain the origin of large scale cosmological magnetic fields (LSMF)~\\cite{giova}. The main difficulty resides in understanding their correlation scale which ranges from the size of galaxies to clusters and super-clusters with an amplitude of the order of micro-gauss, pointing to a primordial origin. Following the work initiated in~\\cite{lat05}, we address this issue in the context of a cold electroweak transition taking place after a period of hybrid inflation. The EW transition has been in the heart of many proposals to address magnetogenesis, linking it in many cases with the generation of the baryon asymmetry. The results presented here resemble the mechanism proposed by Vachaspati connecting the appearance of magnetic fields to that of sphalerons and Z-strings~\\cite{vachas}. The model we have considered is a hybrid inflation model with the bosonic field content of the Standard Model coupled, via the Higgs field, to a singlet inflaton: \\begin{eqnarray} {\\cal L} = - \\frac{1}{4}G^a_{\\mu\\nu}G^{\\mu\\nu}_a - \\frac{1}{4}F^{Y}_{\\mu\\nu}F_{Y}^{\\mu\\nu}+ {\\rm Tr}\\Big[(D_\\mu\\Phi)^\\dag D^\\mu\\Phi\\Big ]+ \\frac{1}{2}(\\partial_\\mu\\chi)^2 - V(\\Phi,\\chi)\\\\ {\\rm V} (\\Phi,\\chi) = {\\rm V}_0 + \\frac{1}{2}(g^2\\chi^2-m^2)\\,|\\Phi|^2 + \\frac{\\lambda}{4} |\\Phi|^4 + \\frac{1}{2} \\mu^2 \\chi^2 \\, \\end{eqnarray} The couplings are fixed to the standard model values for several ratios of the Higgs to W masses. For concreteness, we have fixed the inflaton to Higgs coupling by the relation: $g^2= 2\\lambda $. To solve the time evolution of the system, starting at the end of inflation, we have performed a numerical evolution based on a suitable classical approximation (more details can be found in Refs.~\\cite{marga2,lat05,articulo}). In this work we discuss the results for $\\mh = 4.65\\ \\mw$. Results for more realistic values will be presented in~\\cite{articulo}. ", "conclusions": "We have analysed numerically the proposal that long range magnetic fields could be generated during a cold electroweak transition after a period of low scale hybrid inflation. The generation mechanism is mainly based on two facts: \\begin{itemize} \\item At the SSB stage bubble-like structures, associated to local maxima in the Higgs-field norm, appear. Points outside the bubble front, remaining close to the false vacuum, form string like structures. \\item Bubble collisions give rise to sphaleron-like configurations attached to the location of zeroes of the Higgs field. For $\\theta_W \\ne 0$ these sphalerons behave as magnetic dipoles. \\end{itemize} These two ingredients together lead to an allignment of the sphalerons' dipoles forming magnetic string networks as we have observed in our simulations. Some results concerning the coherence and intensity of the generated magnetic fields have been presented. A further analysis of the time evolution of the magnetic field, as well as the study of the dependence of the $\\mh$ to $\\mw$ ratio will be presented in~\\cite{articulo}. We have also discussed some features of the late time behaviour of the system, among them, thermalisation of photon radiation and turbulence in the scalar fields." }, "0710/0710.2250_arXiv.txt": { "abstract": "Networks of reactions on dust grain surfaces play a crucial role in the chemistry of interstellar clouds, leading to the formation of molecular hydrogen in diffuse clouds as well as various organic molecules in dense molecular clouds. Due to the sub-micron size of the grains and the low flux, the population of reactive species per grain may be very small and strongly fluctuating. Under these conditions rate equations fail and the simulation of surface-reaction networks requires stochastic methods such as the master equation. However, the master equation becomes infeasible for complex networks because the number of equations proliferates exponentially. Here we introduce a method based on moment equations for the simulation of reaction networks on small grains. The number of equations is reduced to just one equation per reactive specie and one equation per reaction. Nevertheless, the method provides accurate results, which are in excellent agreement with the master equation. The method is demonstrated for the methanol network which has been recently shown to be of crucial importance. ", "introduction": "Chemical networks in interstellar clouds consist of gas-phase and grain-surface reactions \\citep{Hartquist1995,Tielens2005}. Reactions that take place on dust grains include the formation of molecular hydrogen \\citep{Gould1963,Hollenbach1971b} as well as reaction networks producing ice mantles and various organic molecules. Unlike gas phase reactions in cold clouds that mainly produce unsaturated molecules, surface processes are dominated by hydrogen-addition reactions that result in saturated, hydrogen-rich molecules, such as H$_2$CO, CH$_3$OH, NH$_3$ and CH$_4$. In particular, recent experiments show that methanol cannot be efficiently produced by gas phase reactions \\citep{Geppert2006}. On the other hand, there are indications that it can be efficiently produced on ice-coated grains \\citep{Watanabe2005}. Therefore, the ability to perform simulations of the production of methanol and other complex molecules on grains is of great importance \\citep{Garrod2006}. Unlike gas-phase reactions, simulated using rate equation models \\citep{Pickles1977,Hasegawa1992}, grain-surface reactions require stochastic methods such as the master equation \\citep{Biham2001,Green2001}, or Monte Carlo (MC) simulations \\citep{Charnley2001}. This is due to the fact that under interstellar conditions, of extremely low gas density and sub-micron grain sizes, surface reaction rates are dominated by fluctuations which cannot be accounted for by rate equations \\citep{Tielens1982,Charnley1997,Caselli1998,Shalabiea1998}. A significant advantage of the master equation over MC simulations is that it consists of differential equations, which can be easily coupled to the rate equations of gas-phase chemistry. Furthermore, unlike MC simulations that require the accumulation of statistical information over long times, the master equation provides the probability distribution from which the reaction rates can be obtained directly. However, the number of equations increases exponentially with the number of reactive species, making the simulation of complex networks infeasible \\citep{Stantcheva2002,Stantcheva2003}. The recently proposed multi-plane method dramatically reduces the number of equations, by breaking the network into a set of fully connected sub-networks \\citep{Lipshtat2004}, enabling the simulation of more complex networks. However, the construction of the multi-plane equations for large networks turns out to be difficult. In this Letter we introduce a method based on moment equations which exhibits crucial advantages over the multi-plane method. The number of equations is further reduced to the smallest possible set of stochastic equations, including one equation for the population size of each reactive specie (represented by a first moment) and one equation for each reaction rate (represented by a second moment). Thus, for typical sparse networks the complexity of the stochastic simulation becomes comparable to that of the rate equations. Unlike the master equation (and the multi-plane method) there is no need to adjust the cutoffs - the same set of equations applies under all physical conditions. Unlike the multi-plane equations, the moment equations are linear and for steady state conditions can be easily solved using algebraic methods. Moreover, for any given network the moment equations can be easily constructed using a diagrammatic approach, which can be automated \\cite{Barzel2007}. ", "conclusions": "In summary, we have introduced a method, based on moment equations, for the simulation of chemical networks taking place on dust-grain surfaces in interstellar clouds. The method provides highly efficient simulations of complex reaction networks under the extreme conditions of low gas density and sub-micron grain sizes, in which the reaction rates are dominated by fluctuations and stochastic simulations are required. The number of equations is reduced to one equation for each reactive specie and one equation for each reaction, which is the lowest possible number for such networks. This method enables us to efficiently simulate networks of any required complexity without compromising the accuracy. It thus becomes possible to incorporate the complete network of surface reactions into gas-grain models of interstellar chemistry. To fully utilize the potential of this method, further laboratory experiments are needed, that will provide the activation energy barriers for diffusion, desorption and reaction processes not only for hydrogen but for all the molecules involved in these networks. We thank A. Lipshtat for helpful discussions. This work was supported by the Israel Science Foundation and the Adler Foundation for Space Research." }, "0710/0710.1409_arXiv.txt": { "abstract": "{% The origin of low-luminosity Type IIP supernovae is unclear: they have been proposed to originate either from massive ($\\sim 25~M_{\\sun}$) or low-mass ($\\sim 9~M_{\\sun}$) stars. }{% We wish to determine parameters of the low-luminosity Type IIP supernova 2003Z, to estimate a mass-loss rate of the presupernova, and to recover a progenitor mass. }{% We compute the hydrodynamic models of the supernova to describe the light curves and the observed expansion velocities. The wind density of the presupernova is estimated using a thin shell model for the interaction with circumstellar matter. }{% We estimate an ejecta mass of $14.0\\pm1.2~M_{\\sun}$, an explosion energy of $(2.45\\pm0.18)\\times10^{50}$ erg, a presupernova radius of $229\\pm39~R_{\\sun}$, and a radioactive $^{56}$Ni amount of $0.0063\\pm0.0006~M_{\\sun}$. The upper limit of the wind density parameter in the presupernova vicinity is $10^{13}$ g\\,cm$^{-1}$, and the mass lost at the red/yellow supergiant stage is $\\leq 0.6~M_{\\sun}$ assuming the constant mass-loss rate. The estimated progenitor mass is in the range of $14.4-17.4~M_{\\sun}$. The presupernova of SN~2003Z was probably a yellow supergiant at the time of the explosion. }{% The progenitor mass of SN~2003Z is lower than those of SN~1987A and SN~1999em, normal Type IIP supernovae, but higher than the lower limit of stars undergoing a core collapse. We propose an observational test based on the circumstellar interaction to discriminate between the massive ($\\sim 25~M_{\\sun}$) and moderate-mass ($\\sim 16~M_{\\sun}$) scenarios. } ", "introduction": "\\label{sec:intro} Type II plateau supernovae (SNe~IIP) with the plateau of $\\sim 100$ days in the light curve are believed to be an outcome of a core collapse of the $9-25~M_{\\sun}$ stars (e.g., Heger et al. \\cite{HFWLH_03}). This paradigm assuming the Salpeter mass spectrum suggests that about 66\\% of all SNe~IIP should be produced by progenitors, i.e., stars on the main sequence, in the $9-15~M_{\\sun}$ range. At present this general picture remains unconfirmed. It could be verified via the determination of ejecta masses from the hydrodynamic modeling for a sufficiently large sample of SNe~IIP. Unfortunately, only few SNe~IIP have the well-observed light curves and spectra needed to reliably reconstruct the basic SN parameters. It is not therefore surprising that up to now ejecta mass has been determined using the detailed hydrodynamic simulations for only SN~1987A and SN~1999em. It is noteworthy that, from the point of view of explosion mechanism, SN~1987A is a normal SN~IIP; it has the ejecta mass, the explosion energy, and the amount of ejected $^{56}$Ni comparable to those of SN~1999em. Recently an interesting subclass of SNe~IIP, so called low-luminosity SNe~IIP, was selected observationally (Pastorello et al. \\cite{PZT_04}). This family is characterized by luminosities, expansion velocities, and radioactive $^{56}$Ni masses which are significantly lower than those of normal SNe~IIP. Two views on the origin of low-luminosity SNe~IIP have been proposed. Turatto et al. (\\cite{TMY_98}) have suggested the origin from massive stars, $\\geq~25~M_{\\sun}$, which presumably form black holes and eject a low amount of $^{56}$Ni; alternatively, these SNe might originate from low-mass stars, $\\sim 9~M_{\\sun}$, which are expected to eject a low amount of $^{56}$Ni (Chugai \\& Utrobin \\cite{CU_00}; Kitaura et al. \\cite{KJH_06}). The investigation of the origin of these SNe~IIP began with some confusion. The point is that SN~1997D, the first low-luminosity SN~IIP, was detected long after the explosion and, therefore, was erroneously claimed to possess a short ($\\sim 50$ days) plateau (Turatto et al. \\cite{TMY_98}). This, in turn, provoked a conclusion that SN~1997D originated from a low-mass ($\\sim 9~M_{\\sun}$) main-sequence star (Chugai \\& Utrobin \\cite{CU_00}). The subsequent discovery of several low-luminosity SNe~IIP with a long plateau of $\\sim 100$ days (Pastorello et al. \\cite{PZT_04}) falsified this conclusion. Until now there were no attempts to model hydrodynamically low-luminosity SNe~IIP on the basis of new observational data. The estimate of the ejecta masses of low-luminosity SN~1999br and SN~2003Z was made only using a semi-analytical model (Zampieri et al. \\cite{ZPT_03}; Zampieri \\cite{Zam_05}). The semi-analytical model, being a sensible tool for the first order estimates, cannot, however, substitute the hydrodynamic simulations. In this paper we study the well-observed low-luminosity Type IIP SN~2003Z (Pastorello \\cite{Pas_03}; Knop et al. \\cite{KHBD_07}). Our approach is based on the hydrodynamic modeling of the light curves and expansion kinematics. This paper is organized as follows. The observational data are presented in Sect.~\\ref{sec:obsdat}. Section~\\ref{sec:model} describes the modeling procedure. Results, specifically, the SN~2003Z parameters and progenitor mass are presented in Sect.~\\ref{sec:results}. The implications of the results are discussed in Sect.~\\ref{sec:discon}. A distance to SN~2003Z of 21.68 Mpc is adopted using the Hubble constant $H_0=70$ km\\,s$^{-1}$\\,Mpc$^{-1}$ and a recession velocity of the host galaxy NGC 2742 $v_{\\mathrm{cor}}=1518$ km\\,s$^{-1}$, corrected for the Local Group infall to the Virgo cluster and taken from the Lyon Extragalactic Data base. There are no observational signatures of the interstellar absorption in the host galaxy (Pastorello \\cite{Pas_03}) so a total extinction is taken to be equal to the Galactic value $A_{B}=0.167$ (Schlegel et al. \\cite{SFD_98}). ", "conclusions": "\\label{sec:discon} The goal of this study was to recover the basic parameters of SN 2003Z and to get an idea about progenitor masses of low-luminosity SNe~IIP. We estimated the ejecta mass to be $14.0\\pm1.2~M_{\\sun}$, the explosion energy $(2.45\\pm0.18)\\times10^{50}$ erg, the pre-SN radius $229\\pm39~R_{\\sun}$, and the $^{56}$Ni mass $0.0063\\pm0.0006~M_{\\sun}$. Using the ejecta/wind interaction model, we found the upper limit of the density parameter of the pre-SN wind and, assuming the constant mass-loss rate at the RSG/YSG stage, constrained the progenitor mass by the range of $14.4-17.4~M_{\\sun}$. The estimate of the wind density thus allows us to avoid the uncertainty in the progenitor mass stemming from the weak sensitivity of the hydrodynamic model to the helium core mass. The only normal SN~IIP, studied in a similar way to SN~2003Z, is SN~1999em. Its pre-SN radius is $\\approx 500~R_{\\sun}$, the ejecta mass is $\\approx 19~M_{\\sun}$, the explosion energy is $\\approx 1.3\\times10^{51}$ erg, and the $^{56}$Ni mass is $\\approx 0.036~M_{\\sun}$ (Utrobin \\cite{Utr_07}), while the progenitor mass is $\\approx 22~M_{\\sun}$ (Chugai et al. \\cite{CCU_07}). A comparison of SN~2003Z with SN~1999em taken together with the fact that the low-luminosity SNe~IIP are very similar indicates that this variety of SNe~IIP originates from less massive progenitors than normal SNe~IIP. Parameters of three SNe~IIP --- SN~1987A (Utrobin \\cite{Utr_05}), SN~1999em (Utrobin \\cite{Utr_07}; Chugai et al. \\cite{CCU_07}), and SN~2003Z (Table \\ref{tab:sumtab}) --- determined on the basis of the similar hydrodynamic modeling, suggest a picture in which ordinary SNe~IIP originate from massive progenitors around $20~M_{\\sun}$, while low-luminosity SNe~IIP originate from less massive progenitors around $\\sim 16~M_{\\sun}$, close to the average mass of the $9-25~M_{\\sun}$ range traditionally associated with SNe~IIP. If our result for SN~2003Z is correct, then the explosion energy and the amount of ejected $^{56}$Ni should significantly increase for the progenitors between $\\sim 16~M_{\\sun}$ and $\\sim 20~M_{\\sun}$ (Fig.~\\ref{fig:nienms}). Interestingly, the empirical correlation between the explosion energy and the $^{56}$Ni mass for normal SNe~IIP has been demonstrated by Nadyozhin (\\cite{Nad_03}). It should be emphasized that these correlations, valid for SNe~IIP, may not be applicable to other SNe~II. At least SN~1994W (Type IIn event) was found to have a low amount of ejected $^{56}$Ni, $\\sim 0.015~M_{\\sun}$, but a normal explosion energy, $\\approx 1.3\\times10^{51}$ erg (Chugai et al. \\cite{CBC_04}). \\begin{table}[t] \\caption[]{Hydrodynamic models for SN 1987A, SN 1999em, and SN~2003Z.} \\label{tab:sumtab} \\centering \\begin{tabular}{@{ } c @{ } c @{ } c @{ } c @{ } c @{ } c @{ } c @{ } c @{ }} \\hline\\hline \\noalign{\\smallskip} SN & $R_0$ & $M_{env}$ & $E$ & $M_{\\mathrm{Ni}}$ & $v_{\\mathrm{Ni}}^{max}$ & $v_{\\mathrm{H}}^{min}$ & $M_\\mathrm{ms}$ \\\\ & $(R_{\\sun})$ & $(M_{\\sun})$ & ($10^{51}$ erg) & $(10^{-2} M_{\\sun})$ & (km\\,s$^{-1}$) & (km\\,s$^{-1}$) & $(M_{\\sun})$ \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 87A & 35 & 18 & 1.5 & 7.65 & 3000 & 600 & 21.5 \\\\ 99em & 500 & 19 & 1.3 & 3.60 & 660 & 700 & 22.2 \\\\ 03Z & 229 & 14 & 0.245 & 0.63 & 535 & 360 & 15.9 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{table} Two alternative conjectures about the origin of low-luminosity SNe~IIP have been proposed: (1) progenitors are massive stars, $\\geq 25~M_{\\sun}$ (Turatto et al. \\cite{TMY_98}); (2) these SNe originate from low-mass stars, $\\sim 9~M_{\\sun}$ (Chugai \\& Utrobin \\cite{CU_00}; Kitaura et al. \\cite{KJH_06}). The present estimate of the progenitor mass of SN~2003Z, $14.4-17.4~M_{\\sun}$, is in the disparity with both the high-mass and low-mass scenarios. Our estimate of the progenitor mass is essentially based on the assumption that the mass-loss rate at the RSG/YSG stage was constant. In fact, the derived wind density refers to the close vicinity of SN~2003Z $r10$ km\\,s$^{-1}$ corresponds to the wind history during the latest $R_{\\mathrm w}/u <300$ yr before the explosion. Most of the RSG/YSG stage, i.e. $4\\times10^5-10^6$ yr, is thus hidden from our sight. One might suggest that the mass-loss rate was substantially more vigorous in the past than immediately before the SN explosion. If this questionable possibility were the case, the $25~M_{\\sun}$ progenitor would lose $\\sim 10~M_{\\sun}$, and the high-mass scenario for SN~2003Z would be saved. Of note, the essentially higher wind density of pre-SN in this case suggests that the dense wind at the large radii could be revealed via radio and X-ray observations at the large age, $t\\geq10$ yr. A remarkable property of all three SNe is strong mixing between the helium core and the hydrogen envelope indicated by a low minimal velocity of hydrogen matter (Table \\ref{tab:sumtab}). Generally, the model with the unmixed helium core shows a bump in the light curve at the end of the plateau (Utrobin \\cite{Utr_07}); the absence of the bump in the available light curves of SNe~IIP, in addition to the narrow-topped H$\\alpha$ emission at the nebular phase, indicates that substantial mixing between the helium core and the hydrogen envelope is a universal phenomenon for SNe~IIP. This mixing could be induced by the convection during the growth of helium core (Barkat \\& Wheeler \\cite{BW_88}) or/and during the SN explosion (Kifonidis et al. \\cite{KPSJM_06}). \\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics{fig9.eps}} \\caption{% Explosion energy (\\textbf{a}) and $^{56}$Ni mass (\\textbf{b}) versus progenitor mass for three core-collapse SNe. } \\label{fig:nienms} \\end{figure} It is noteworthy that the radii of all three pre-SNe~IIP (Table \\ref{tab:sumtab}) are notably lower than a radius of typical massive RSG, e.g., the radius of $\\approx 800~R_{\\sun}$ for $\\alpha$ Ori (Harper et al. \\cite{HBL_01}). We have already shown that the SN~2003Z pre-SN was the YSG rather than the RSG. For SN~1999em the $22~M_{\\sun}$ pre-SN with the luminosity of $\\approx 10^5~L_{\\sun}$ is characterized by the effective temperature of $\\sim 4700$~K, which is intermediate between the RSG and YSG values. Interestingly, the pre-SN of SN~2004et, a normal SN~IIP, was probably a YSG as well (Li et al. \\cite{LVFC_05}). These cases suggest that not only the pre-SN~1987A, but a possibly significant fraction of pre-SNe~IIP, experience an excursion on the Hertzsprung-Russell (HR) diagram from the RSG towards the YSG before the explosion. Note, some pre-SNe~IIP, when becoming the YSG, could get into the instability strip and manifest themselves before the explosion as long period ($P\\sim50-100$ days) Cepheids. If this is the case, the recovery of the Cepheid period of a pre-SN on the basis of a set of pre-explosion observations could provide us with an additional tool for the mass determination from the pulsation period. The recovered progenitor mass of SN~2003Z raises a crucial question: what happens to the $9-15~M_{\\sun}$ stars which make up $\\approx66$\\% of all the stars from the $9-25~M_{\\sun}$ mass range. Two conceivable suggestions are: (A) all the $9-15~M_{\\sun}$ stars explode as low-luminosity, or even fainter, SNe~IIP; (B) low-luminosity SNe~IIP originate only from a narrow range of progenitors around $\\sim 16~M_{\\sun}$, while the rest of the $9-15~M_{\\sun}$ stars produce different varieties of SNe~II. A large sample of SNe~II which is well observed and studied is certainly needed to distinguish between the options A and B. It should be emphasized that the initial peak, the end of the plateau, and the radioactive tail are of crucial importance for the recovery of reliable SN~II parameters. An alternative approach to determine the pre-SN mass exploits archival pre-explosion images of host galaxies and stellar evolution tracks on the HR diagram. Interestingly, using the archival HST images Van Dyk et al. (\\cite{VLF_03}) and Maund \\& Smartt (\\cite{MS_05}) derived an upper limit of $\\sim 15~M_{\\sun}$ and $\\sim 12~M_{\\sun}$, respectively, for the progenitor of the low-luminosity SN~1999br. Given a similarity of known low-luminosity SNe~IIP these estimates make the high-mass scenario unlikely. It should be emphasized that a test of the mass determination method based on the pre-explosion images is needed; for instance, the pre-SN light could be partially absorbed in a dusty circumstellar envelope. In this regard it would be of top priority to determine independently the progenitor mass both from the pre-discovery images and from the hydrodynamic simulations. Among SNe~IIP, except for SN~1987A, only SN~2004et and SN~2005cs were detected in the pre-discovery images (Li et al. \\cite{LVFC_05}; Maund et al. \\cite{MSD_05}; Li et al. \\cite{LVF_06}) and have the light curves and spectra appropriate for the hydrodynamic modeling as well (Sahu et al. \\cite{SASM_06}; Misra et al. \\cite{MPC_07}; Pastorello et al. \\cite{PST_06}; Tsvetkov et al. \\cite{TVS_06}). These two challenging cases require a detailed study." }, "0710/0710.1123_arXiv.txt": { "abstract": "Optical and near-infrared spectroscopy of the newly discovered peculiar L dwarf {\\name} are presented. Folkes et al.\\ identified this source as a high proper motion L9$\\pm$1 dwarf based on its strong H$_2$O absorption at 1.4~$\\micron$. We find that the optical spectrum of {\\namesh} is in fact consistent with that of a normal L4.5 dwarf with notably enhanced FeH absorption at 9896~{\\AA}. However, its near-infrared spectrum is unusually blue, with strong H$_2$O and weak CO bands similar in character to several recently identified ``blue L dwarfs''. Using {\\namesh} as a case study, and guided by trends in the condensate cloud models of Burrows et al.\\ and Marley et al., we find that the observed spectral peculiarities of these sources can be adequately explained by the presence of thin and/or large-grained condensate clouds as compared to normal field L dwarfs. Atypical surface gravities or metallicities alone cannot reproduce the observed peculiarities, although they may be partly responsible for the unusual condensate properties. We also rule out unresolved multiplicity as a cause for the spectral peculiarities of {\\namesh}. Our analysis is supported by examination of {\\em Spitzer} mid-infrared spectral data from Cushing et al.\\ which show that bluer L dwarfs tend to have weaker 10~$\\micron$ absorption, a feature tentatively associated with silicate oxide grains. With their unique spectral properties, blue L dwarfs like {\\namesh} should prove useful in studying the formation and properties of condensates and condensate clouds in low temperature atmospheres. ", "introduction": "L dwarfs comprise one of the two latest-type spectral classes of very low mass stars and brown dwarfs, spanning masses at and below the hydrogen burning minimum mass (see \\citealt{kir05} and references therein). They are inexorably linked to the presence and properties of liquid and solid condensates which form in their cool photospheres (e.g., \\citealt{tsu96,bur99,ack01,all01,coo03,tsu05}). These condensates significantly influence the spectral energy distributions and photospheric gas abundances of L dwarfs, by removing gaseous TiO and VO from the photosphere and enabling the retention of atomic alkali species (e.g., \\citealt{feg96,bur99,lod02}). Weakened {\\wat} absorption through backwarming effects (e.g., \\citealt{jon97,all01}) and red near-infrared colors ($J-K$ $\\approx$ 1.5--2.5; \\citealt{kir00}) also result from condensate opacity. In addition, periodic and aperiodic photometric variability observed in several L dwarfs has been associated with surface patchiness in photospheric condensate clouds (e.g., \\citealt{bai99,bai01,gel02,moh02}). Condensate abundances at the photosphere appear to reach their zenith amongst the mid- and late-type L dwarfs \\citep{kir99,cha00,ack01} before disappearing from the photospheres of cooler T dwarfs \\citep{mar96,tsu96b,all01,cus06}. The abundances of photospheric condensates, their grain size distribution, and the radial and surface structure of condensate clouds may vary considerably from source to source, as well as temporally for any one source, and the dependencies of these variations on various physical parameters are only beginning to be explored \\citep{hel01,woi03,kna04}. With hundreds of L dwarfs now known,\\footnote{A current list is maintained at \\url{http://dwarfarchives.org}.} groupings of peculiar L dwarfs -- sources whose spectral properties diverge consistently from the majority of field objects -- are becoming distinguishable. Examples include young, low surface gravity brown dwarfs \\citep{mcg04,kir06,all07,cru07} and metal-poor L subdwarfs \\citep{me0532,lep1610,giz06,megmos}. There also exists a class of peculiar ``blue'' L dwarfs \\citep{cru03,cru07,kna04}, roughly a dozen sources exhibiting normal optical spectra but unusually blue near-infrared colors and strong near-infrared {\\wat}, FeH and {\\ki} features. Various studies have attributed these peculiarities to subsolar metallicity, high surface gravity, unresolved multiplicity and peculiar cloud properties \\citep{giz00,cru03,cru07,mcl03,mewide3,kna04,chi06,fol07}. Any one of these characteristics may impact the presence and character of condensates and condensate clouds in low temperature atmospheres. In an effort to identify new nearby and peculiar L dwarfs, we have been searching for late-type dwarfs using near-infrared imaging data from the Deep Near Infrared Survey of the Southern Sky (DENIS; \\citealt{epc97}). One of the objects identified in this program is DENIS~J112639.9$-$500355, a bright source which was concurrently discovered by \\citet{fol07} in the SuperCOSMOS Sky Survey \\citep[hereafter SSS]{ham01a,ham01b,ham01c} and the Two Micron All Sky Survey (hereafter 2MASS; \\citealt{skr06}). It is designated {\\name} in that study, and we refer to the source hereafter as {\\namesh}. Based on its blue near-infrared colors and deep {\\water} absorption bands, \\citet{fol07} concluded that {\\namesh} is a very late-type L dwarf (L9$\\pm$1) which may have unusually patchy or thin condensate clouds. In this article, we critically examine the observational properties of {\\namesh} to unravel the origins of its spectral peculiarities, and examine it as a representative of the blue L dwarf subgroup. Our identification of {\\namesh} and a slightly revised determination of its proper motion using astrometry from the SSS, DENIS and 2MASS catalogs are described in $\\S$~2. Optical and near-infrared spectroscopic observations and their results are described in $\\S$~3, along with determination of the optical and near-infrared classifications of {\\namesh} and estimates of its distance and space kinematics. In $\\S$~4 we analyze the properties of {\\namesh} and blue L dwarfs in general, considering metallicity, surface gravity, condensate cloud and unresolved multiplicity effects. We also introduce a new near-infrared {\\wat} index that eliminates discrepancies between optical and near-infrared types for these sources. Results are discussed and summarized in $\\S$~5. ", "conclusions": "Our analysis in $\\S$~4.2 leads us to conclude that the spectral peculiarities of {\\namesh} and other blue L dwarfs have their immediate cause in condensate cloud effects, specifically the presence of thin, patchy or large-grained condensate clouds at the photosphere. Subsolar metallicities and high surface gravities in of themselves cannot reproduce the observed spectral peculiarities of these sources. However, it is clear that these latter physical properties must play a role in determining the cloud characteristics of blue L dwarfs. Lower metallicities reduce the metal species available to form condensates, resulting in less condensate material overall. Higher surface gravities may increase the sedimentation rate of condensate grains, potentially resulting in thinner clouds. The large tangential velocities and absence of {\\lii} absorption in the three blue L dwarfs {\\namesh}, 2MASS~J1300+1921 and 2MASS~J1721+3344 support the idea that these sources may be relatively old and possibly slightly metal-poor. However, the influence of other physical parameters on condensate cloud properties must also be considered, including rotation rates, vertical upwelling rates (e.g., \\citealt{sau06}) and possibly magnetic field strengths. An assessment of how these fundamental physical parameters influence the properties of condensate clouds in low-temperature atmospheres is the subject of ongoing theoretical investigations (e.g., \\citealt{hel01,woi03}; M.\\ Marley, in preparation). Empirical studies are also necessary, particularly those focused on well-characterized samples of blue (and red) L dwarfs. To this end, Table~\\ref{tab_blue} lists all blue L dwarfs currently reported in the literature. We anticipate that this list will grow as near-infrared spectroscopic follow-up of L dwarfs continues." }, "0710/0710.1065_arXiv.txt": { "abstract": "We use a semi-analytic circumstellar disk model that considers movement of the snow line through evolution of accretion and the central star to investigate how gas giant frequency changes with stellar mass. The snow line distance changes weakly with stellar mass; thus giant planets form over a wide range of spectral types. The probability that a given star has at least one gas giant increases linearly with stellar mass from 0.4\\,$M_\\odot$ to 3\\,$M_\\odot$. Stars more massive than 3\\,$M_\\odot$ evolve quickly to the main-sequence, which pushes the snow line to 10--15\\,AU before protoplanets form and limits the range of disk masses that form giant planet cores. If the frequency of gas giants around solar-mass stars is 6\\%, we predict occurrence rates of 1\\% for 0.4\\,$M_\\odot$ stars and 10\\% for 1.5\\,$M_\\odot$ stars. This result is largely insensitive to our assumed model parameters. Finally, the movement of the snow line as stars $\\gtrsim$2.5\\,$M_\\odot$ move to the main-sequence may allow the ocean planets suggested by \\citeauthor{2004Icar..169..499L} to form without migration. ", "introduction": "\\label{sec:intro} In the last ten years, the discovery of more than 200 extra-solar planets,\\footnote{http://vo.obspm.fr/exoplanetes/encyclo/encycl.html} and more than 200 debris disks,\\footnote{http://www.roe.ac.uk/ukatc/research/topics/dust/identification.html } suggests that planet formation is a common and robust process. Planet masses inferred from debris disks range from terrestrial to Jovian, at distances as great as tens of AU from the central star \\citep[e.g.][]{2004AJ....127..513K,2005ApJ...619L.187G}. The nature and sensitivity of radial velocity surveys means that most of the planets are $\\sim$Jupiter mass gas giants in close orbits around Sun-like stars. However, recent discoveries as diverse as icy $\\sim$Neptune-mass planets orbiting M dwarfs \\citep[e.g.][]{2005ApJ...634..625R}, and debris disks around A-type stars \\citep[e.g.][]{2005ApJ...620.1010R} show that planet formation occurs over a wide range of spectral types. Current theory suggests that planets form in similar ways around all stars. Thus, the increasing diversity of stellar hosts and planetary systems provides an opportunity to test these theories. For this reason, the types of planets most likely to form around stars of differing spectral types has become a renewed area of study \\citep[e.g.][]{2005ApJ...626.1045I,2006ApJ...643..501B,2006A&A...458..661K,2006ApJ...650L.139K}, after the idea was first explored by \\citeauthor{1988MNRAS.235..193N} nearly 20 years ago \\citep{1987MNRAS.224..107N,1988MNRAS.230..551N,1988MNRAS.235..193N}. Theories of Solar System formation generally include the ``snow line,'' where ices condense from the nebular gas. The snow line distance is usually fixed in a disk with a time independent surface density and temperature profile around a main-sequence star \\citep[e.g.][]{2005ApJ...626.1045I}. In a more realistic picture, the disk and stellar properties evolve considerably during the 1--10\\,Myr pre--main-sequence (PMS) lifetime when planets probably form \\citep[e.g.][]{1987Icar...69..249L,1996Icar..124...62P}. As the disk temperature evolves with time, movement of the snow line may therefore influence the properties of theoretical planetary systems \\citep[e.g.][]{2006ApJ...650L.139K,2007ApJ...654..606G}. Here, we begin to develop a time dependent model for the formation of gas giant cores that considers the PMS evolution of the star and surrounding accretion disk. We introduce a simple semi-analytic disk model, based on the ``minimum mass solar nebula,'' that links movement of the snow line through evolution of disk accretion and stellar luminosity. In contrast to previous studies \\citep[e.g.][]{2005ApJ...626.1045I,2006A&A...458..661K}, our analysis suggests that gas giant formation around stars more massive than the sun is more likely than around less massive stars. We cover the background important to our story in \\S \\ref{sec:background}, consider the snow line in \\S \\ref{sec:motivation}, and outline our model in \\S \\ref{sec:model}. We present our results in \\S \\ref{sec:results}, and discuss and conclude in \\S \\ref{sec:discussion} and \\S \\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} We describe a model for the evolution of the snow line in a planet forming disk, and apply it over a range of stellar masses to derive the probability distribution of gas giants as a function of stellar mass. The two main ingredients for our model are a prescription for movement of the snow line due to accretion and PMS evolution, and rules that determine whether protoplanets are massive enough, and form early enough, to become gas giants. The snow line distance generally moves inward over time. With our prescription for the accretion rate, accretion dominates over irradiation for stars with $M_\\star \\lesssim 2\\,M_\\odot$. For $\\gtrsim$3\\,$M_\\odot$ stars, irradiation dominates at times $\\gtrsim$1\\,Myr as the star moves up to its main-sequence luminosity. The transition is at a few Myr for $\\sim$2\\,$M_\\odot$ stars. Over the wide range of observed accretion rates for any fixed stellar mass, the snow line in some disks may be set entirely by irradiation. The snow line generally sets where the innermost gas giant cores form. In relatively massive disks around intermediate mass stars, rocky cores form interior to the snow line. The location of the outermost core is always set by the gas dissipation timescale. The range of disk masses that form cores, and the radial width of the region in the disk where they form, increase with stellar mass. Lower mass disks produce failed icy cores, which are probably similar to Uranus, Neptune, and the observed ``super-Earths.'' The outward movement of the snow line as stars more massive than the Sun reach the main-sequence, and as the disk becomes optically thin, allows the ocean planets suggested by \\citet{2004Icar..169..499L} to form \\emph{in situ}. The change in disk temperature is only large enough for these planets to harbour oceans around stars $\\gtrsim$2.5\\,$M_\\odot$. Our model includes several poorly determined parameters, which current and future facilities will investigate. While there are current resolved studies of gaseous disks \\citep[e.g.][]{2007arXiv0704.1481B}, the next generation of telescopes such as GMT and ALMA will provide more information on surface density profiles, and how disk properties change with stellar mass and age. These studies will help to constrain input parameters for our model. We have shown that the time dependence of the snow line in part determines where gas giant cores form. This result should motivate future studies of planet formation in disks whose properties change with time. The subsequent evolution of isolated cores is beyond the scope of this paper, but further work that investigates the growth and dynamical evolution of these objects can investigate the diversity of resulting system structures. Given an initial distribution of disk masses, the probability that a star has at least one gas giant increases linearly with stellar mass from 0.4\\,$M_\\odot$ to 3\\,$M_\\odot$. If the frequency of gas giants around solar-mass stars is 6\\%, we predict an occurrence rate of 1\\% (10\\%) for 0.4\\,$M_\\odot$ (1.5\\,$M_\\odot$) stars. This result is largely insensitive to changes in our model parameters. In contrast to the \\citet{2005ApJ...626.1045I} model, where it is hard to form observable gas giants above 1\\,$M_\\odot$, our model predicts a peak at $\\sim$3\\,$M_\\odot$ because we include disk and PMS evolution in our snow line derivation. However, our model does not include migration, so our prediction applies to observable and currently undetectable gas giants. Though sample numbers are small, it appears that observable gas giant frequency increases with stellar mass across a wide range of host masses \\citep{2007arXiv0707.2409J}. Larger samples of stars that host giant planets, particularly low and intermediate-mass stars, will solidify this result. These studies, and the extension of the results to a wider range of semi-major axes, will provide a basis for comparison with our model predictions." }, "0710/0710.5271_arXiv.txt": { "abstract": "The unveiled main-sequence splitting in $\\omega$ Centauri as well as NGC 2808 suggests that matter highly-enriched in He (in terms of its mass fraction $Y\\sim0.4$) was produced and made the color of some main-sequence stars bluer in these globular clusters (GCs). The potential production site for the He-rich matter is generally considered to be massive AGB stars that experience the second dredge-up. However, it is found that massive AGB stars provide the matter with $Y\\sim 0.35$ at most, while the observed blue-shift requires the presence of $Y\\sim 0.4$ matter. Here, we show that extra mixing, which operates in the red giant phase of stars less massive than $\\sim2\\,M_{\\odot}$, could be a mechanism that enhances He content in their envelopes up to $Y\\sim 0.4$. The extra mixing is supposed to be induced by red giant encounters with other stars in a collisional system like GCs. The $Y\\sim 0.4$ matter released in the AGB phase has alternative fates to (i) escape from a GC or (ii) be captured by kinematically cool stars through encounters. The AGB ejecta in $\\omega$ Cen, which follows the latter case, can supply sufficient He to cause the observed blue-shift. Simultaneously, this scheme generates the extreme horizontal branch, as observed in $\\omega$ Cen in response to the higher mass loss rates, which is also caused by stellar encounters. ", "introduction": "The discovery of a split in the main sequence (MS) of \\omcen\\ \\citep{Bedin04} has opened a new window in the study of Galactic globular clusters (GCs). The observed fact that stars of the blue MS (bMS) in \\omcen\\ exhibit slightly stronger absorption of iron lines on average, in comparison with the red MS (rMS) \\citep{Piotto05} which strongly implies that the origin of bMS is attributable to the enhancement of He inside bMS stars. Subsequently, the recent HST observation has provided us with the second sample of MS splitting, NGC 2808 \\citep{Piotto07}. This GC essentially exhibits no dispersion of [Fe/H] and thus the argument of He enhancement in bMS stars has been reinforced. A comparison between the observed color-magnitude diagrams (CMDs) and synthetic population models suggests that bMS He abundance is enhanced to $Y\\sim 0.4$ for both of two GCs \\citep{Norris04, Lee05, DAntona05, Piotto07}. \\citet{Tsujimoto07} have argued that if the surface of a star on the MS is polluted with the $Y=0.4$ matter with the mass of $\\sim 0.1 \\msun$, the star can move to a position on the observed bMS in the CMD. This relaxes a severe demand on the amount of He possibly supplied from asymptotic giant branch (AGB) stars, in contrast to the idea that the bMS stars are born from pure AGB ejecta consisting of $Y\\sim 0.4$. Similarly, \\citet{Newsham07} have insisted that surface pollution may explain the high He content of bMS stars. It should be stressed that $Y\\sim 0.4$ matter is necessary to realize the bMS stars in these GCs. Since massive AGB stars can enhance the He abundance in their envelopes through the second dredge-up in the early AGB phase, they have been considered a possible production site for the $Y\\sim 0.4$ matter \\citep{DAntona05}. However, previous studies indicated that the resultant abundance of He in the envelope of any AGB star does not exceed $Y \\sim 0.35$ \\citep[see e.g.,][]{vandenHoek97}. Therefore, such a He yield from massive AGB stars may not be sufficient to split the MS as observed. In the subsequent section, we will discuss this issue and conclude that it is highly implausible for a massive AGB star to enhance He abundance up to $Y\\sim 0.4$ in its envelope. Then, we should search for an alternative mechanism that enhances the He abundance more efficiently than the second dredge-up. We believe that this process, which enhances He abundance, should be accompanied by some other elemental signatures. This is reminiscent of the abundance anomaly of red giants in GCs. The abundance anomalies observed for red giants, such as the abundance variations of CNO elements and the O-Na anticorrelation, are common attributes associated with GCs including \\omcen\\ \\citep[see][]{Gratton04}. One of the proposed mechanisms to explain these anomalies is a deep mixing, so-called extra mixing, on the red giant stage \\citep{Sweigart79, Langer93, Charbonnel98, Fujimoto99,Aikawa01,Chaname05,Suda06}. The extra mixing is assumed to be caused by some kind of rotation-induced mixing, driven by shear instability around the border of the He core \\citep[see][]{Fujimoto88,Zahn92}. Since such a deep mixing naturally draws He from the core, the extra mixing is a promising mechanism of producing He-rich matter. As a mechanism of mixing, most of the previous works assumed a continuous mixing caused by the meridional circulation, while Fujimoto and his coworkers insisted that a flash of H brought into a degenerate core by the rotation-induced mixing drives the intermittent mixing. In this study, we propose that extra mixing, which occurs in the red giant phase of a star, enhances He abundance in the envelope and ejects He-rich matter when the star evolves to a white dwarf through mass loss. In this scenario, the driving force of the extra mixing is generated through interactions between the red giant and dwarfs in the central region of \\omcen. Our detailed calculations reveal that stars with the masses less than $\\sim$2 \\msun\\ can enhance He abundance up to $Y\\sim 0.4$, provided that the extra mixing operates in the red giant phase. Based on our scenario, we discuss other conspicuous characteristics of \\omcen, the extremely blue horizontal branch (HB), which is observed in some other Galactic GCs. Although factors that determine HB morphology is still an open question and is known as the second parameter problem, the most effective parameter that affects the location of a HB star in the CMD is the mass $M_{\\rm env}$ of the envelope. A major factor to determine $M_{\\rm env}$ is the mass loss during the red giant phase. We predict that encounters induce a high mass loss rate through a gain of angular momentum and promote the presence of an extremely blue HB population. In the end, two mysteries in \\omcen\\ of MS splitting and HB morphology are discussed in a unified framework of encounters of dwarfs with AGB ejecta and those of red giants with dwarfs. ", "conclusions": "We first propose that He-rich matter is synthesized in red giants with the masses of 0.8 - 2 $\\msun$, which experience encounters with other stars in GCs, leading to the extra mixing during the RGB phase in their envelopes. On the other hand, massive AGB stars are unable to produce the He abundance as much as $Y \\sim 0.4$ because (1) their large envelopes work as a buffer against He enrichment and (2) the amount of dredged-up matter is limited by thinning the He-burning shell, due to the growth of C-O core. Therefore, a $Y\\sim 0.4$ matter is a product unique to GCs. In some GCs, this matter is retained in their gravitational potential for a prolonged time and is accreted by kinematically cool stars, although this may not be the case for most GCs. In addition, stellar encounters in GCs induce the extra mixing and increase the mass loss rate during the RGB phase. As a result, such stars are predicted to evolve to significantly blue HB stars. In conclusion, MS splitting is inclined to be accompanied by the existence of an extremely blue HB, as observed in \\omcen\\ and NGC 2808. It should be, however, noted that our models might make a variation in the final He abundance as a result of extra mixing induced by various amounts of angular momentum transfered through encounters. If the variation is too large, then it will erase the MS splitting and lead to a broad MS. Further investigations of the extra mixing model in terms of the MS splitting are surely awaited." }, "0710/0710.0662_arXiv.txt": { "abstract": "The study of Wolf-Rayet stars plays an important role in evolutionary theories of massive stars. Among these objects, $\\sim 20\\%$ are known to be in binary systems and can therefore be used for the mass determination of these stars. Most of these systems are not spatially resolved and spectral lines can be used to constrain the orbital parameters. However, part of the emission may originate in the interaction zone between the stellar winds, modifying the line profiles and thus challenging us to use different models to interpret them. In this work, we analyzed the He{\\sevensize II}$\\lambda$4686\\AA \\ + C{\\sevensize IV}$\\lambda$4658\\AA \\ blended lines of WR30a (WO4+O5) assuming that part of the emission originate in the wind-wind interaction zone. In fact, this line presents a quiescent base profile, attributed to the WO wind, and a superposed excess, which varies with the orbital phase along the 4.6 day period. Under these assumptions, we were able to fit the excess spectral line profile and central velocity for all phases, except for the longest wavelengths, where a spectral line with constant velocity seems to be present. The fit parameters provide the eccentricity and inclination of the binary orbit, from which it is possible to constrain the stellar masses. ", "introduction": "Massive stars are known to drive strong winds, which are responsible for the transfer of a large amount of the stellar mass to the interstellar medium, contributing to the feedback of chemical elements, and to the creation of cloud cavities in which these objects are found. Typically, O stars present mass-loss rates of $\\dot{M}_{\\rm O} \\sim 10^{-6} - 10^{-5}$ M$_\\odot$ yr$^{-1}$ and wind velocities of $v_{\\rm O} \\sim 2000 - 3500$ km s$^{-1}$, while Wolf-Rayet (WR) stars present $\\dot{M}_{\\rm WR} \\sim 10^{-6} - 10^{-4}$ M$_\\odot$ yr$^{-1}$ and $v_{\\rm WR} \\sim 1000 - 4000$ km s$^{-1}$ (Nugis \\& Lamers 2000, Lamers 2001). In massive binary systems in which both stars present high mass-loss rates and high-velocity winds, the collision of the winds will occur. Therefore, a contact surface is formed where the momenta of the two winds are equal, surrounded by two shocks. The post-shocked gas, cools as it flows along the contact surface and is responsible for strong free-free emission at X-ray and radio wavelengths. X-rays in massive binary systems are orders of magnitude higher than those observed in single massive stars, and are used as an indication of binarity. UV and optical lines can also indicate the binary nature of a given object, when they present periodic profile variations. If the lines are of photospheric or atmospheric origin, as the stars move along their orbit, they suffer Doppler shifts, which depend on the orbital phase and inclination. However, some massive objects show periodic variable line profiles that cannot be explained under these assumptions. Seggewiss (1974) noted that the binary system WR79 presented two peaks, superimposed to the C{\\sevensize III} emission line, which changed their position and intensity with time. Typically, in double line spectroscopic binaries, each peak moves in a different direction, indicating opposite velocity components along the line of sight for each star; however in WR79 both peaks moved in the same direction. L$\\rm \\ddot{u}$hrs (1997) presented a model in which the two peaks were not produced by the stellar photosphere, but were generated by the flowing gas at the contact surface between the two strong shocks. This model reproduced well the data for WR 79, but failed to reproduce the line profiles of other WR binary systems, among them WR 30a (Bartzakos, Moffat \\& Niemela 2001). Falceta-Gon\\c calves, Abraham \\& Jatenco-Pereira (2006) improved L$\\rm \\ddot{u}$hrs' model introducing more realistic parameters, as stream turbulence and gas opacity, to account for the line broadening and peak displacement. In the present work, we applied this model to WR 30a, which was classified as a WO4+O5 binary system (Moffat \\& Seggewiss 1984, Crowther et al. 1998); its binary nature was reported by Niemela (1995) based on spectral-line radial velocities obtained with a high temporal resolution. Later, Gosset et al. (2001) presented a detailed analysis of the spectra of WR30a for several epochs and were able to confirm the binary hypothesis and to determine the period of $P \\sim 4.6$ days. They also noted strong line-profile variations, which made more difficult the determination of the stellar mass ratio and the orbital inclination from standard methods. They concluded that the C{\\sevensize IV}$\\lambda$4658\\AA \\ line-profile variations were related to wind-wind collision processes, but did not model the lines under such an assumption. The same conclusion was reached by Bartzakos, Moffat \\& Niemela (2001) using the C{\\sevensize IV}$\\lambda$5801\\AA \\ line. Also, Paardekooper et al. (2003) presented photometric measurements at $V$ and $B$ bands; the light curves confirmed the period obtained by Gosset et al. (2001), but also showed higher frequency variability in the $V$ band, with timescale of hours. They concluded that this could be due to the strong variability of the C{\\sevensize IV}$\\lambda$5801\\AA \\ line, possibly related to the wind-wind interaction. In the present work, we tested the wind-wind shock emission hypothesis on the C{\\sevensize IV}$\\lambda$4658 \\AA \\ excess line profile variations measured by Gosset et al. (2001) using the model developed by Falceta-Gon\\c calves, Abraham \\& Jatenco-Pereira (2006), which is briefly described in Section 2. In Section 3, we show the results obtained for WR 30a and present a brief discussion, followed by the conclusions in Section 4. ", "conclusions": "In this work we presented an application of the wind-wind shock emission model proposed by Falceta-Gon\\c calves et al. (2006) to the line profile variations of WR 30a. In this model, the wind-wind shock structure can be represented by a cone, along which the shocked material flows. This gas will emit radiation, cool down and eventually reach recombination temperatures. The observed emission lines will then suffer Doppler shifts due to the stream velocity component along the line of sight. During the orbital movement, the cone position will change, as well as the radial velocity, which will cause the line profile variations observed in several massive binary systems. Gosset et al. (2001) obtained detailed spectra of WR30a during more than 30 days. They determined the orbital period ($P = 4.6$d), and obtained the radial velocity curve for the O-star. Regarding the WR component, they found that the blended He{\\sevensize II}$\\lambda$4686\\AA \\ and C{\\sevensize IV}$\\lambda$4658\\AA \\ lines showed a variable excess emission. In the present paper we modeled this variable emission, being able to reproduce the variations except for the red part of the profiles, which seemed to be unchanged in velocity and were probably generated in the stellar wind of the WR star instead of in the shock region. The best-fitting result was obtained for $\\beta = 50^\\circ$, $i = 20^\\circ$, $\\sigma = 0.3$ and $v_{\\rm flow} = 2200$km s$^{-1}$. Also, correlating the orbital phase with the modeled phase angle, it was possible to determine the orbital eccentricity as $e = 0.2$, similar to the value of 0.0 previously assumed {\\it ad hoc} by other authors. Although both values lead to very small differences in the orbital shape, its value is important for the determination of the stellar masses and orbital separation between the stars. Using this eccentricity and orbital inclination to model the radial velocity curve for the O-star, we found $K_{\\rm O} = 25$ km s$^{-1}$. If we assume $M_{\\rm O} = 40 - 60$ M$_{\\odot}$, we find $M_{\\rm WR} = 7.5 - 9.7$ M$_{\\odot}$, and the orbital major semi-axis $a_{\\rm O} \\simeq 5.4$ R$_{\\odot}$ and $a_{\\rm WR} \\simeq 30$ R$_{\\odot}$. The model showed itself a powerful tool for constraining the wind and orbital parameters of massive binary systems. The fits did not matched the data exactly at all epochs, but considering the difficulties of subtracting the emission excess from the original spectra, the general shape and peak position variations were well reproduced. We must state that this is a very simple approximation, taking into account the complexities found in such systems. Numerical simulations could give a more detailed analysis, and probably more accurate values for the model parameters in future works." }, "0710/0710.0381_arXiv.txt": { "abstract": "The evolution of the galaxy stellar mass--star formation rate relationship ($M_*-$SFR) provides key constraints on the stellar mass assembly histories of galaxies. For star-forming galaxies, $M_*-$SFR is observed to be fairly tight with a slope close to unity from $z\\sim 0\\rightarrow 2$, and it evolves downwards roughly independently of $M_*$. Simulations of galaxy formation reproduce these trends, broadly independent of modeling details, owing to the generic dominance of smooth and steady cold accretion in these systems. In contrast, the observed amplitude of the $M_*-$SFR relation evolves markedly differently than in models, indicating either that stellar mass assembly is poorly understood or that observations have been misinterpreted. Stated in terms of a star formation activity parameter $\\asf\\equiv (M_*/SFR)/(t_{\\rm Hubble}-{\\rm 1 Gyr})$, models predict a constant $\\asf\\sim 1$ out to redshifts $z\\sim 4+$, while the observed $M_*-$SFR relation indicates that $\\asf$ increases by $\\sim\\times 3$ from $z\\sim 2$ until today. The low $\\asf$ (i.e. rapid star formation) at high-$z$ not only conflicts with models, but is also difficult to reconcile with other observations of high-$z$ galaxies, such as the small scatter in $M_*-$SFR, the slow evolution of star forming galaxies at $z\\sim 2-4$, and the modest passive fractions in mass-selected samples. Systematic biases could significantly affect measurements of $M_*$ and SFR, but detailed considerations suggest that none are obvious candidates to reconcile the discrepancy. A speculative solution is considered in which the stellar initial mass function (IMF) evolves towards more high-mass star formation at earlier epochs. Following Larson (1998), a model is investigated in which the characteristic mass $\\hat{M}$ where the IMF turns over increases with redshift. Population synthesis models are used to show that the observed and predicted $M_*-$SFR evolution may be brought into general agreement if $\\hat{M}=0.5(1+z)^{2} M_\\odot$ out to $z\\sim 2$. Such IMF evolution matches recent observations of cosmic stellar mass growth, and the resulting $z=0$ cumulative IMF is similar to the ``paunchy\" IMF favored by \\citet{far07} to reconcile the observed cosmic star formation history with present-day fossil light measures. ", "introduction": "How galaxies build up their stellar mass is a central question in galaxy formation. Within the broadly successful cold dark matter scenario, gas is believed to accrete gravitationally into growing dark matter halos, and then radiative processes enable the gas to decouple from non-baryonic matter and eventually settle into a star-forming disk~\\citep[e.g.][]{mo98}. However, many complications arise when testing this scenario against observations, as the stellar assembly of galaxies involves a host of other processes including star formation, kinetic and thermal feedback from various sources, and merger-induced activity, all of which are poorly understood in comparison to halo assembly. A key insight into gas accretion processes is the recent recognition of the importance of ``cold mode\" accretion~\\citep{kat03,bir03,ker05}, where gas infalling from the intergalactic medium (IGM) does not pass through an accretion shock on its way to forming stars. Simulations and analytic models show that cold mode dominates global accretion at $z\\ga 2$~\\citep{ker05}, and dominates accretion in all halos with masses $\\la 10^{12}$M$_\\odot$~\\citep{bir03,ker05}. The central features of cold mode accretion are that it is (1) {\\it rapid} (2) {\\it smooth}, and (3) {\\it steady}. It is rapid because it is limited by the free-fall time and not the cooling time; it is smooth because most of the accretion occurs in small lumps and not through major mergers~\\citep{mur01,ker05,guo07}; and it is steady because it is governed by the gravitational potential of the slowly growing halo. A consequence of cold mode accretion is that the star formation rate is a fairly steady function of time~\\citep[e.g.][]{fin06,fin07}, which results in a tight relationship between the stellar mass $M_*$ and the star formation rate SFR. Hence simulations of star-forming galaxies generically predict a tight relationship with $M_*\\propto$SFR that evolves slowly with redshift. In detail, a slope slightly below unity occurs owing to the growth of hot halos around higher-mass galaxies that retards accretion. The stellar mass--star formation rate ($M_*-$SFR) relation for star forming galaxies has now been observed out to $z\\sim 2$, thanks to improving multiwavelength surveys. As expected from models, the relationship is seen to be fairly tight from $z\\sim 0-2$, with a slope just below unity and a scatter of $\\la 0.3$~dex~\\citep{noe07a,elb07,dad07}. It also evolves downwards in amplitude to lower redshift in lock step fashion, suggesting a quiescent global quenching mechanism such as a lowering of the ambient cosmic density, again as expected in models. The range of data used to quantify this evolution is impressive: \\citet{noe07a} used the AEGIS multi-wavelength survey in the Extended Groth Strip to quantify the $M_*$-SFR relation from $z\\sim 1\\rightarrow 0$; \\citet{elb07} used GOODS at $z\\sim 0.8-1.2$ and SDSS spectra at $z\\sim 0$; and \\citet{dad07} used star-forming BzK-selected galaxies with {\\it Spitzer}, X-ray, and radio follow-up to study the relation at $z\\sim 1.4-2.5$. In each case, a careful accounting was done of both direct UV and re-radiated infrared photons to measure the total galaxy SFR, paying particular attention to contamination from active galactic nuclei (AGN) emission (discussed in \\S\\ref{sec:syst}). Multiwavelength data covering the rest-optical were employed to accurately estimate $M_*$. This broad agreement in $M_*-$SFR slope, scatter, and qualitative evolution between model predictions and these data lends support to the idea that cold mode accretion dominates in these galaxies. Yet all is not well for theory. Closer inspection reveals that the {\\it amplitude} of the $M_*-$SFR relation evolves with time in a way that is inconsistent with model expectations. This disagreement is fairly generic, as shown in \\S\\ref{sec:comparison}, arising in both hydrodynamic simulations and semi-analytic models, and is fairly insensitive to feedback implementation. The sense of the disagreement is that going to higher redshifts, the observed SFRs are higher, and/or stellar masses lower, than predicted in current models. The purpose of this paper is to understand the implications of the observed $M_*-$SFR relation for our theoretical view of how galaxies accumulate stellar mass. In \\S\\ref{sec:tact} it is argued that $M_*-$SFR amplitude evolution implies a typical galaxy star formation history that is difficult to reconcile with not only model expectations, but also other observations of high-$z$ galaxies. Possible systematic effects that may bias the estimation of SFR and $M_*$ are considered in \\S\\ref{sec:syst}, and it is argued that none of them are obvious candidates to explain the discrepancy. Finally a speculative avenue for reconciliation is considered, namely that the stellar initial mass function (IMF) has a characteristic mass that evolves with redshift (\\S\\ref{sec:imf}). This model is constrained based on the $M_*$-SFR relation in \\S\\ref{sec:imfmodel}. The implications for such an evolving IMF are discussed in \\S\\ref{sec:mchar}. Results are summarized in \\S\\ref{sec:summary}. A $\\Lambda$CDM cosmology with $\\Omega=0.25$ and $H_0=70$~km/s/Mpc is assumed throughout for computing cosmic timescales. ", "conclusions": "\\label{sec:summary} Implications of the observed stellar mass--star formation rate correlation are investigated in the context of current theories for stellar mass assembly. The key point, found here and pointed out in previous studies~\\citep{dad07,elb07}, is that the amplitude of the $M_*-$SFR relation evolves much more rapidly since $z\\sim 2$ in observations that in current galaxy formation models. It is shown here that this is true of both hydrodynamic simulations and semi-analytic models, and is broadly independent of feedback parameters. In contrast, the slope and scatter of the observed $M_*-$SFR relation are in good agreement with models. The tight $M_*-$SFR relation with a slope near unity predicted in models is a generic result owing to the dominance of cold mode accretion, particularly in early galaxies, which produces rapid, smooth and relatively steady infall. The slow amplitude evolution arises because star formation starts at $z\\ga 6$ for moderately massive star-forming galaxies and continues at a fairly constant or mildly declining level to low-$z$. The large discrepancy in amplitude evolution when compared to observations from $z\\sim 2\\rightarrow 0$ hints at some underlying problem either with the models or with the interpretation of data. A convenient parameterization of the problem is through the star formation activity parameter, $\\asf\\equiv (M_*/{\\rm SFR})/(t_H-1\\;{\\rm Gyr})$, where $t_H$ is the Hubble time. A low value corresponds to a starbursting system, a high value to a passive system, and a value near unity to system forming stars constantly for nearly a Hubble time. In models, $\\asf$ remains constant around unity from $z=0-4$, whereas in observations it rises steadily from $\\la 0.2$ at $z\\sim 2$ to close to unity at $z\\sim 0$. Several ad hoc modifications to the theoretical picture of stellar mass assembly are considered in order to match $\\asf$ evolution, but each one is found to be in conflict with other observations of high-redshift galaxies. Bursts seem unlikely given the low scatter in $M_*-$SFR. Delaying galaxy formation to match $\\asf$ results in such a low redshift for the onset of star formation, it is in conflict with observations of star forming galaxies at earlier epochs. Hiding a large population of galaxies as passive runs into difficulty when compared to direct observations of the passive galaxy fraction in mass-selected samples at $z\\sim 2$. Having an exponentially growing phase of star formation or ``staged\" galaxy formation are perhaps the most plausible solutions from an observational viewpoint, but it is difficult to understand theoretically how such scenarios can arise within hierarchical structure formation. Hence if observed star-forming galaxies are quiescently forming stars and represent the majority of galaxies at $z\\sim 2$, as other observations seem to suggest, then it is not easy to accomodate the low $\\asf$ inferred from the $M_*-$SFR amplitude. Systematic uncertainties in $M_*$ or SFR determinations could bias results progressively more at high-$z$ in order to mimic evolution in $\\asf$. Various currently debated sources are considered, such as the contribution to near-IR light from TP-AGB stars, extinction corrections, assumptions about star formation history, AGN contamination, and PAH emission calibration. It may be possible to concoct scenarios whereby several of these effects combine to mimic $\\asf$ evolution, but there are no suggestions from local observations that such scenarios are to be expected. Hence if systematic effects are to explain the low $\\asf$ at high-$z$, it would imply significant and unexpected changes in tracers of star formation and stellar mass between now and high redshifts. A solution is proposed that the stellar initial mass function becomes increasingly bottom-light to higher redshifts. Several lines of arguments are presented that vaguely or circumstantially favor such evolution, though no smoking gun signatures are currently known. A simple model of IMF evolution is constructed, based on the ansatz by \\citet{lar98} that the minimum temperature of molecular clouds is reflected in the characteristic mass of star formation ($\\hat{M}_{\\rm IMF}$) where the mass contribution per logarithmic mass bin is maximized. The minimum temperature may increase with redshift owing to a hotter cosmic microwave background, more vigorous star formation activity, or lower metallicities within early galactic ISMs. In order to reconcile the observed $\\asf$ evolution with theoretical expectations of no evolution, an evolving IMF of the form \\begin{eqnarray*} \\frac{dN}{d\\log{M}} &\\propto& M^{-0.3}\\;\\; {\\rm for}\\; M<\\hat{M}_{\\rm IMF},\\\\ &\\propto& M^{-1.3}\\;\\; {\\rm for}\\; M>\\hat{M}_{\\rm IMF};\\\\ \\hat{M}_{\\rm IMF}&=&0.5 (1+z)^{2} M_\\odot \\end{eqnarray*} is proposed. The exponent of $\\hat{M}_{\\rm IMF}$ evolution is constrained by requiring no $\\asf$ evolution, through careful modeling with the PEGASE.2 population synthesis code. While the exact form of IMF evolution is not well constrained by present observations, what is required is that the IMF has progressively more high-mass stars compared to low-mass at earlier epochs, and that this IMF applies to the majority of star forming galaxies at any epoch. It is worth noting that this evolving IMF is only constrained out to $z\\sim 2$ from the $M_*-$SFR relation, though it yields predictions that are consistent with other observations out to $z\\sim 4$. Extrapolating such evolution to higher redshifts is dangerous, since no observational constraints exist and the precise cause of IMF evolution is not understood. Implications of such an evolving IMF are investigated. By leaving the high-mass end of the IMF unchanged, recent successes in understanding the connections between high-mass star formation, feedback, and metal enrichment are broadly preserved. The cosmic stellar mass accumulation rate would be altered compared to what is inferred from cosmic star formation history measurements using a standard IMF. It is shown that an evolving IMF is at face value in better agreement with direct measures of cosmic stellar mass assembly~\\citep{per07,els07,wil07}. Furthermore, the evolving IMF goes towards relieving the generic tension between present-day fossil light measures versus observations of the cosmic star formation history. In particular, the paunchy IMF favored by \\citet{far07} in order to reconcile the observed cosmic star formation history, present-day $K$-band luminosity density, and extragalactic background light constraints, is qualitatively similar to the cumulative IMF of all stars formed by today in the evolving IMF case. Individually, each argument has sufficient uncertainties to cast doubt on whether a radical solution such as an evolving IMF is necessary. But taken together, the $M_*-$SFR relation adds to a growing body of circumstantial evidence that the ratio of high-mass to low-mass stars formed is higher at earlier epochs. It is by no means clear that IMF variations are the only viable solution to the $M_*-$SFR dilemma. The claim here is only that an evolving IMF is an equally (un)likely solution as invoking unknown systematic effects or carefully crafted star formation histories in order to explain $M_*-$SFR evolution. It is hoped that this work will spur further efforts, both observational and theoretical, to investigate this important issue. Future plans include performing a more careful analysis of the evolving IMF in terms of observable quantities, in order to accurately quantify the impact of such IMF evolution on the interpretation of UV, near-IR, and mid-IR light. Also, it is feasible to incorporate such an evolving IMF directly into simulation runs to properly account for gas recycling in such a scenario. Finally, including some feedback mechanism to truncate star formation in massive systems may impact the $M_*-$SFR relation in some way, so this will be incorporated into the hydro simulations. Observationally, pushing SFR and $M_*$ determinations to higher redshifts is key; a continued drop in $\\asf$ to $z\\sim 3$ would rapidly solidify the discrepancies with current models, and would also generate stronger conflicts with other observations of high-$z$ galaxies. Assessing the AGN contribution and extinction uncertainties from high-$z$ systems is critical for accurately quantifying the light from high-mass star formation, for instance through the use of more direct star formation indicators such as Paschen-$\\alpha$. Pushing observations further into the mid-IR such as to 70$\\mu$, past the PAH bands at $z\\sim 2$, would mitigate PAH calibration uncertainties in SFR estimates. Obtaining a large sample of spectra for typical high-$z$ star-forming systems \\citep[like cB58;][]{pet00} would more accurately constrain the SED than broad-band data. All of these programs push current technological capabilities to their limits and perhaps beyond, but are being planned as facilities continue their rapid improvement. In summary, owing to the robust form of star formation histories in current galaxy formation models, the $M_*$-SFR relation represents a key test of our understanding of stellar mass assembly. Current models reproduce the observed slope and scatter remarkably well, but broadly fail this test in terms of amplitude evolution. Whether this reflects some fundamental lack of physical insight, or else some missing ingredient such as an evolving IMF, is an issue whose resolution will have a significant impact on our understanding of galaxy formation." }, "0710/0710.5047_arXiv.txt": { "abstract": "{We present results of an analysis of the optical spectrum of the post-AGB star HD\\,56126 (IRAS\\,07134\\,+\\,1005) based on observations made with the echelle spectrographs of the 6-m telescope with spectral resolutions of R\\,=\\,25000 and 60000 at 4012--8790\\,\\AA. The profiles of strongest lines (HI; FeII, YII, BaII absorptions, etc.) formed in the expanding atmosphere at the base of the stellar wind have complex and variable shapes. To study the kinematics of the atmosphere, the velocities of individual features in these profiles must be measured. Differential line shifts of up to $\\rm V_r$\\,=\\,15$\\div$30\\,km/s have been detected from the lines of metals and molecular features. The star's atmosphere simultaneously contains both expanding layers and layers falling onto the star. A comparison of the data for different times demonstrates that both the radial velocity and the velocity pattern in whole are variable. The position of the molecular spectrum is stable, implying stability of the expansion velocity of the circumstellar envelope around HD\\,56126 detected in observations in the C$_2$ and NaI lines.} \\authorrunning{Klochkova \\& Chentsov} \\titlerunning{HD\\,56126: structure of the atmosphere and envelope} ", "introduction": "We studied the kinematic parameters of the atmosphere of the star HD\\,56126 (SAO\\,96709), which belongs to the asymptotic giant branch (the post-AGB stage). In this short evolution stage, stars are in the process of their transition from the AGB to becoming a planetary nebula; for this reason, they are generally known as proto-planetary nebulae (PPNe). In the Hertzsprung--Russell diagram, post-AGB stars evolve toward the left from the AGB, maintaining nearly constant luminosity while becoming hotter. As descendants of AGB stars, these objects can be used to trace variations in the physical conditions and chemical parameters of the stellar material due to changes in the sources of energy release in the stellar interior, the ejection of the envelope, and mixing. HD\\,56126 is the optical component of the IR source IRAS\\,07134+1005, which has a double peaked spectral energy distribution (SED), typical for PPNe. In addition to this anomalous SED, associated with the circumstellar dust envelope, the star also displays other characteristics of this type of object [\\cite{Kwok}]: the optical component of the PPN is a F5\\,Iab supergiant outside the galactic plane (b=+9$\\lefteqn{.}^m$99); the central star is surrounded by an extended nebula whose angular size exceeds 4$^{\\rm ''}$, according to Hubble Space Telescope observations [\\cite{Ueta}] (the largest known for this type of PPN); and the optical spectrum displays H$\\alpha$ emission and absorption with a variable line profile [\\cite{Oud1994}]. In addition to these general PPN characteristics noted by Kwok [\\cite{Kwok}], subsequent studies of HD\\,56126 and the associated IR source revealed several important peculiarities expected for this evolution stage: a large excess of carbon and $s$-process elements [\\cite{Kloch1995}] and an emission feature at $\\lambda$\\,=\\,21\\,$\\mu$ in the IR--spectrum. In the small subgroup of PPN having this emission, the presence of this 21\\,$\\mu$ feature was found to correlate with observational manifestations of the products of nucleosynthesis (the excess of carbon and heavy metals in the outer layers of the atmosphere) [\\cite{Kloch1998, Decin}]. Thus, HD\\,56126 displays all the properties ex pected for a post-AGB object, indicating the importance of detailed studies of its optical spectrum with high spectral resolution in a broad wavelength range. Fortunately, HD\\,56126 is fairly bright (B\\,=\\,9$\\lefteqn{.}^m$11, V = 8.27m), making it the most convenient carbon enriched PPN star for high resolution spectroscopy. The main moments of our study are spectral features identification and comparison the spectra of HD\\,56126 and the standard $\\alpha$\\,Per (Sp\\,=\\,F5\\,Iab); search for profile variability for spectral features; analyses the radial velocities $\\rm V_r$ to search for differential shifts; and studying the variability of the radial velocity. In Section~2, we briefly describe the techniques used for the observations, processing, and analysis of the spectral data. The main conclusions concerning the spectrum of the star are presented in Section~3.1, while Sections~3.2 and 3.3 present our radial velocity measurements derived using various features of the spectrum, and discuss the temporal behavior of the radial-velocity pattern. Our results are summarized in Section~4. {\\footnotesize \\begin{table}[t] \\caption{Moments of observations and values of heliocentric radial velocitity $\\rm V_r$ measured. Column~4 contains $\\rm V_r$ values averaged over weak lines (with depths R$_\\lambda$ close to the continuum level, R$_\\lambda \\rightarrow$0). Velocities corresponding to the positions of the strongest components are presented for FeII(42), H$\\alpha$, and D2\\,NaI, with values determined from the weakest components given in parantheses. The two velocities in italics in column~5 are determined from the IR--oxygen triplet OI\\,7773\\,\\AA. Uncertain values are marked by a colon.} \\medskip \\begin{tabular}{ll c| c c c l l l @{\\quad} | l l l} \\hline \\small Date &\\small Spectro-- &\\small $\\Delta\\lambda$ & \\multicolumn{7}{c}{\\small $V_r$} \\\\ \\cline{4-12} &\\small graph &\\small \\AA{} & R$_\\lambda \\rightarrow $0 & \\small Fe{\\sc ii} &\\small H$\\beta$ & \\quad \\small H$\\alpha$ & \\small D\\,Na{\\sc i}& \\small C$_2$ &\\multicolumn{3}{c}{\\small intestellar} \\\\ \\hline \\quad 1& 2& 3&4&5&6&\\quad7& 8&9& 10& 11 & 12\\\\ \\hline 12.01.93&Lynx&5560--8790& 88.8 &{\\it 91} & &78 (100:)&77 &79: & & & \\\\ 10.03.93&Lynx&5560--8790& 89.0 &{\\it 93} & &71 (43:) &75: &76: & & & \\\\ 04.03.99&Lynx&5050--6640& 85.9 & 77 & &76 (43:) &78 &77.1& & & \\\\ 20.11.02&NES &4560--5995& 89.6 &95 (80:) &89 & &74.9 (89) &77.2&12.0 &23.5 & 30.8\\\\ 21.02.03&NES &5150--6660& 88.8 & 96: & &88 (112:)&75.6 (89) &77.1&12 &24 & 31 \\\\ 12.04.03&NES &5270--6760& 88.4 & & &82 (103:)&75.4 (89:)& &13 &23 & 30.5 \\\\ 14.11.03&NES &4518--6000& 85.3 &96 (86:) &97 & &75.0 (87:)&76.9&12.5 & & \\\\ 10.01.04&NES &5270--6760& 86.7 & & &54: (78:)&75.6 (86:)& &13.0 &23.5 & 31 \\\\ 09.03.04&NES &5275--6767& 89.8 & & &58 (74:) &76.1 (89) & &13 &24 & 31 \\\\ 12.11.05&NES &4010--5460& 82.5 &97 (77:) &98 & & &77.5& & & \\\\ \\hline \\end{tabular} \\end{table} } ", "conclusions": "Our high resolution spectra of HD\\,56126 have revealed variability of various line profiles: along with the previously known H$\\alpha$ profile variability, the profiles of strongest lines (such as BaII, YII, and FeII) also proved to be variable. The broad spectral range encompassed by our spectra enabled us to measure radial velocities using spectral features that form at various depths in the stellar atmosphere and circumstellar envelope. We were able to distinguish the C$_2$ absorption molecular bands (and their structure), as well as bands identified with diffuse interstellar bands. We found substantial differential shifts in lines of different intensities within the same spectrum, i.e., appreciable $\\rm V_r$(R) dependences, as well as variation of these shifts with time. This indicates the need for velocity measurements based on a large set of lines in an extended spectral interval. We conclude that both expanding layers and matter falling onto the star exist simultaneously in the stellar atmosphere. The position of the molecular spectrum is stable, indicating stability of the expansion velocity of the envelope." }, "0710/0710.2451_arXiv.txt": { "abstract": "Recently the numerical simulations of the process of reionization of the universe at $z>6$ have made a qualitative leap forward, reaching sufficient sizes and dynamic range to determine the characteristic scales of this process. This allowed making the first realistic predictions for a variety of observational signatures. We discuss recent results from large-scale radiative transfer and structure formation simulations on the observability of high-redshift Ly-$\\alpha$ sources. We also briefly discuss the dependence of the characteristic scales and topology of the ionized and neutral patches on the reionization parameters. ", "introduction": "The observations of high-redshift QSO's \\citep{2001AJ....122.2833F,2001AJ....122.2850B} and large-scale CMB polarization \\citep{2007ApJS..170..377S} indicate that the intergalactic medium has been completely ionized by redshift $z\\sim6$ through an extended process. The most probable cause was the ionizing radiation of the First Stars and QSO's. Currently these are the two main direct observational constraints on this epoch. This scarcity of observational data is set to change dramatically in the next few years, however. A number of large observational projects are currently under way, e.g. observations at the redshifted 21-cm line of hydrogen \\citep[e.g.][]{1997ApJ...475..429M,2000ApJ...528..597T, 2002ApJ...572L.123I,2006MNRAS.372..679M,2006PhR...433..181F}, detection of small-scale CMB anisotropies due to the kinetic Sunyaev-Zel'dovich (kSZ) effect \\citep[e.g.][]{2000ApJ...529...12H,2003ApJ...598..756S,kSZ}, and surveys of high-redshift Ly-$\\alpha$ emitters and studies of the IGM absorption \\citep[e.g.][]{2003AJ....125.1006R,2004ApJ...604L..13S, 2006NewAR..50...94B}. The planning and success of these experiments relies critically upon understanding the large-scale geometry of reionization, i.e. the size- and spatial distribution of the ionized and neutral patches. This is best derived by large-scale simulations, although a number of semi-analytical models exist as well \\citep[e.g.][]{2004ApJ...613....1F}. Recently we presented the first large-scale, high-resolution radiative transfer simulations of cosmic reionization \\citep{2006MNRAS.369.1625I,% 2007MNRAS.376..534I} and applied those to derive a range of reionization observables \\citep{2006MNRAS.372..679M,kSZ,pol21,cmbpol,wmap3,2007arXiv0708.3846I}. Here we summarize recent results on the characteristic scales and topology of reionization and implications of our simulations for the observability of high-redshift Ly-$\\alpha$ sources. ", "conclusions": "" }, "0710/0710.4936_arXiv.txt": { "abstract": "We interpret the recent gravitational lensing observations of Jee et al. \\cite{Jee} as first evidence for a {\\it caustic} ring of dark matter in a galaxy cluster. A caustic ring unavoidably forms when a cold collisionless flow falls with net overall rotation in and out of a gravitational potential well. Evidence for caustic rings of dark matter was previously found in the Milky Way and other isolated spiral galaxies. We argue that galaxy clusters have at least one and possibly two or three caustic rings. We calculate the column density profile of a caustic ring in a cluster and show that it is consistent with the observations of Jee et al. ", "introduction": "Using strong and weak gravitational lensing methods, Jee et al. \\cite{Jee} constructed a column density map of the central region of the galaxy cluster Cl 0024+1654. The map shows a ring of dark matter of radius $\\simeq$ 400 kpc, width $\\sim$ 150 kpc and maximum column density $\\simeq 58 {M_\\odot \\over {\\rm pc}^2}$. Remarkably, the overdensity in dark matter is not accompanied by an analogous structure in x-ray emitting gas or luminous matter. In this regard, Jee et al. discovered a new instance where dark and ordinary matter have dramatically different spatial distributions. The well-known observations of the ``Bullet Cluster\" 1E0657-56 \\cite{bul} provided an earlier example. Jee et al. interpret the dark matter ring in Cl 0024+1654 as the product of a near head-on collision, along the line of sight, of two subclusters. They performed a simulation of the response of the dark matter particles to the time-varying gravitational field and found that, after the collision has occurred, the dark matter particles move outward and form shell-like structures which appear as a ring when projected along the collision axis \\cite{Jee}. The interpretation of Jee et al. fits with independent lines of evidence that Cl 0024+1654, an apparently relaxed cluster, is a collision of two subclusters. However, it is shown in ref. \\cite{ZuH} that the observed dark matter ring is reproduced only for highly fine-tuned, and hence unlikely, initial velocity distributions. The purpose of our paper is to propose an alternative interpretation, to wit that Jee et al. have observed a caustic ring formed by the in and out flow of dark matter particles falling onto the cluster for the first time. We show that the observed ring is explained assuming only that the dark matter falling onto the cluster has net overall rotation, with angular momentum vector close to the line of sight, and velocity dispersion less than 60 km/s. When cold collisionless dark matter falls from all directions into a smooth gravitational potential well, the phase space distribution of the dark matter particles is characterized everywhere by a set of discrete flows \\cite{Ips}. The flows form outer and inner caustics. The outer caustics are formed by outflows where they turn around before falling back in. Each outer caustic is a fold catastrophe ($A_2$) located on a topological sphere surrounding the potential well. The inner caustics \\cite{crdm,sing} are formed near where the particles with the most angular momentum in a given inflow reach their closest approach to the center before going back out. The catastrophe structure of the inner caustics \\cite{inner} depends on the angular momentum distribution of the infalling particles. If that angular momentum distribution is characterized by net overall rotation, the inner caustics are rings (closed tubes) whose cross-section is a section of the elliptic umbilic catastrophe ($D_{-4}$) \\cite{sing}. These statements are valid independently of any assumptions of symmetry, self-similarity, or anything else. It is argued in ref.~\\cite{rob} that discrete flows and caustics are a generic and robust property of {\\it galactic} halos if the dark matter is collisionless and cold. The radii of the outer caustic spheres are predicted by the self-similar infall model \\cite{FG,B} of halo formation. The radii of the inner caustic rings are predicted \\cite{crdm}, in terms of a single parameter $j_{\\rm max}$, after the model is generalized \\cite{STW} to allow angular momentum for the infalling particles. Evidence for inner caustic rings distributed according to the predictions of the self-similar infall model has been found in the Milky Way \\cite{crdm,milk,mon} and in other isolated spiral galaxies \\cite{crdm,Kinn}. The resolution of most present numerical simulations is inadequate to see discrete flows and caustics. However such features are seen in dedicated simulations which increase the number of particles in the relevant regions of phase space \\cite{Stiff,simca}. They should also become apparent in fully general simulations of structure formation through the use of special techniques \\cite{White}. ", "conclusions": "We conclude that the dark matter ring observed by Jee et al. has properties consistent with a caustic ring of dark matter. The column density agrees both in shape and overall amplitude. At present the data are too imprecise to infer with confidence the nature of the observed ring. However, our proposal makes a distinct prediction for the column density profile across the ring, as illustrated in Fig.1. This may be tested by future observations. In this regard, let us emphasize that the ${1 \\over x} I_a({2(x-a) \\over s})$ profile applies to each azimuth. If the caustic ring interpretation is confirmed, an important corollary is that dark matter falls onto Cl 0024+1654 with net overall rotation. ~~ We thank Igor Tkachev for making available his numerical codes to solve the equations of the self-similar infall model. We thank Leanne Duffy for useful discussions. This work was supported in part by the U.S. Department of Energy under contract DE-FG02-97ER41029. P.S. gratefully acknowledges the hospitality of the Aspen Center of Physics while working on this project." }, "0710/0710.3727_arXiv.txt": { "abstract": "The energetics and emission mechanism of GRBs are not well understood. Here we demonstrate that the instantaneous peak flux or equivalent isotropic peak luminosity, $L_{iso}$ ergs s$^{-1}$, rather than the integrated fluence or equivalent isotropic energy, $E_{iso}$ ergs, underpins the known high-energy correlations. Using new spectral/temporal parameters calculated for 101 bursts with redshifts from {\\em BATSE}, {\\em BeppoSAX}, {\\em HETE-II} and {\\em Swift} we describe a parameter space which characterises the apparently diverse properties of the prompt emission. We show that a source frame characteristic-photon-energy/peak luminosity ratio, $K_{z}$, can be constructed which is constant within a factor of 2 for all bursts whatever their duration, spectrum, luminosity and the instrumentation used to detect them. The new parameterization embodies the Amati relation but indicates that some correlation between $E_{peak}$ and $E_{iso}$ follows as a direct mathematical inference from the Band function and that a simple transformation of $E_{iso}$ to $L_{iso}$ yields a universal high energy correlation for GRBs. The existence of $K_{z}$ indicates that the mechanism responsible for the prompt emission from all GRBs is probably predominantly thermal. ", "introduction": "The energetics of the central engine which powers the explosion responsible for a GRB are both intriguing and fundamental to our understanding of these cosmic events. The isotropic energy outflow at source, estimated using the integrated gamma-ray fluence, is enormous, up to $E_{iso}\\sim10^{54}$ ergs, and even if the outflow is collimated in jets the total energy involved is still huge, $E_{\\gamma}\\sim10^{51}$ ergs. The possibility that the explosion taps a standard energy resevoir has been pursued by many authors following the initial suggestion from Frail et al. (2001). If this total energy available were, indeed, roughly constant (or predictable through other means) and we could reliably estimate the collimation, then GRBs could be used as a cosmological probe to very high redshifts, Bloom et al. (2003), Ghirlanda et al. (2004). Early on it was noted that, based on analysis of {\\em BATSE} data, there was a correlation between $E_{p}$, the peak of $E.F(E)$ where $F(E)$ ergs cm$^{-2}$ keV$^{-1}$ is the observed spectrum, and the fluence (Mallozzi et al. 1995, Lloyd et al. 2000). When redshifts became available for long bursts the isotropic energy, $E_{iso}$, could be estimated from the fluence and the peak energy could be transformed into the source frame, $E_{pz}$, the so-called Amati relation, a correlation between $E_{iso}$ and $E_{pz}$ in the sense that more energetic bursts have a higher $E_{pz}$, was discovered using data from {\\em BeppoSAX}, (Amati et al. 2002). This correlation has subsequently been confirmed and extended although there remain many significant outliers, including all short bursts. The physical origin of the correlation may be associated with the emission mechanisms operating in the fireball but the theoretical details are far from settled (see the discussion by Amati (2006) and references therein). More recently a tighter correlation between $E_{iso}$, $E_{pz}$ and the jet break time, $t_{break}$, measured in the optical afterglow has been reported (Ghirlanda et al. 2004). This is explained in terms of a modification to the Amati relation in which $E_{iso}$ is corrected to a true collimated energy, $E_{\\gamma}$, using an estimate of the collimation angle derived from $t_{break}$. The details of the collimation correction depend on the density and density profile of the circumburst medium, Nava et al. (2006) and references therein. Multivariable regression analysis was performed by Liang \\& Zhang (2005) to derive a model-independent relationship, $E_{iso}\\propto E_{pz}^{1.94}t_{zbreak}^{-1.24}$, indicating that the rest-frame break time of the optical afterglow, $t_{zbreak}$ was indeed correlated with the prompt emission parameters. Other studies have concentrated on the properties of the isotropic peak (maximum) luminosity, $L_{iso}$ ergs s$^{-1}$, measured over some short time scale $\\approx1$ s, rather than the time integrated isotropic energy, $E_{iso}$. Yonetoku et al. (2004) noted a correlation between $L_{iso}$ and $E_{pz}$ for 16 GRBs with firm redshifts. A correlation between $L_{iso}$ and the spectral lag was first identified by Norris et al. (2000) and explained in terms of the evolution of $E_{peak}$ with time. The shocked material responsible for the gamma-ray emission is expected to cool at a rate proportional to the gamma-ray luminosity and it has been suggested that $E_{peak}$ traces the cooling (Schaefer 2004). A similar correlation between $L_{iso}$ and the variability of the GRB ($V$) was described by Reichart et al. (2001). The origin of the $L_{iso}-V$ relation is likely to be related to the physics of the relativistic shocks and the bulk Lorentz factor of the outflow. It could be that high $\\Gamma_{outflow}$ results in high $L_{iso}$ and $V$ while lower luminosity and variability are expected if $\\Gamma_{outflow}$ is low (see, for example, M\\'{e}sz\\'{a}ros et al. 2002). A rather bizzare correlation involving $L_{iso}$, $E_{pz}$ and variability was found by Firmani et al. (2006). They employed the ``high signal'' time, $T_{45}$, as formulated by Reichart et al. (2001) in their study of variability, and showed that $L_{iso}\\propto E_{pz}^{1.62}T_{45}^{-0.49}$ for 19 GRBs with a spread much narrower than that of the Amati relation. There is currently no explanation for such a correlation although it may be connected with the spectral lag and variability correlations and the Amati relation. The correlation between $E_{iso}$ and $E_{pz}$ supplemented by additional empirical information can be used in pseudo redshift indicators, for example Atteia (2003), Pelangeon \\& Atteia (2006), but the intrinsic spread in the correlation and uncertainty about the underlying physical interpretation introduce errors, typically of a factor $\\sim2$. It may be possible to reduce the errors by simultaneous application of several independent luminosity/energy correlations, and extension of the Hubble Diagram to high redshifts using GRBs has been attempted, see for example Schaefer (2007). However, it is not clear that the correlations briefly described above are truly independent and there may be some underlying principle or mechanism which connects them all together. Recently, and more controversially, Butler et al. (2007) have raised serious doubts about the validity of these correlations suggesting that it is likely that they are introduced by observational/instrumental bias and have nothing to do with the physical properties of the GRBs and hence they conclude that GRBs are probably useless as cosmological probes. Here we take a new look at the source frame spectral and temporal properties of a large number of GRBs for which we have redshifts in order to try and understand what really correlates with what and whether or not this can provide useful intrinsic information about the GRBs and what drives them. In this analysis we include the short-duration GRBs which may share a similar emission mechanism with long bursts despite probably having different progenitors. ", "conclusions": "The equivalent isotropic energy, $E_{iso}$ ergs, of a GRB can be expressed as the product of two source frame terms, a characteristic photon energy, $E_{wz}$ keV, calculated from the shape of the spectrum across the range 1-10000 keV and the energy density at the peak of the $E.F_{z}(E)$ spectrum, $Q_{pz}$ ergs keV$^{-1}$. The correlation trend between $E_{wz}$ and $Q_{pz}$ gives rise to the Amati relation. By stacking the samples of a GRB light curve into descending order we can construct a rate profile. The functional form of such rate profiles is common to the vast majority of bursts. Fitting the profile gives us a luminosity time, $T_{Lz}$ s, a measure of the burst duration which can be used to convert the energy density at the peak to a luminosity density at peak, $Q_{pz}/T_{Lz}$ ergs keV$^{-1}$ s$^{-1}$. We can calculate the peak equivalent isotropic luminosity as a product $L_{iso}=E_{wz}Q_{pz}/T_{Lz}=E_{iso}/T_{Lz}$ ergs s$^{-1}$. $E_{wz}$ is a characteristic photon energy or a measure of the colour or hardness of the burst and $Q_{pz}/T_{Lz}$ is a measure of the instantaneous peak brightness. We have gathered and analysed sufficient spectral and temporal data from 101 bursts to produce the relation between $E_{wz}$ vs. $Q_{pz}/T_{Lz}$ and $E_{wz}$ vs. $L_{iso}$, shown in Figure \\ref{fig11}, which constitutes the closest thing we have to an intrinsic colour-magnitude diagram for the peak emission from GRBs, $E_{wz}\\propto L_{iso}^{0.25}$. All bursts are clustered such that we can construct a intrinsic colour-magnitude quasi constant $K_{z}$, which is a function of the source frame characteristic photon energy/peak luminosity ratio given by Equation \\ref{eq15}. The range of equivalent isotropic energy that drives the expanding fireball is very large, 6 orders of magnitude (Figure \\ref{fig3}), but the instantaneous hardness/brightness of the peak emission covers a very small intrinsic dynamic range, $\\approx4$. The existence and form of $K_{z}$ indicates that the physical mechanism for the Gamma-ray production at the photosphere of the fireball is common to all bursts and is probably thermal although many other possibilities are not ruled out. If the prompt spectra are dominated by thermal photons the scatter in $K_{z}$ may be attributed to variations in the size and/or Lorentz factor of the fireball. XRFs have low $\\Gamma_{0}$ and/or large radii. Short bursts have high $\\Gamma_{0}$ and/or small radii. The relation between $T_{Lz}$ vs. $Q_{pz}$ clearly separates short from long, but both classes have the same instantaneous peak hardness/brightness." }, "0710/0710.5667_arXiv.txt": { "abstract": "We present a review of the standard paradigm for giant planet formation, the core accretion theory. After an overview of the basic concepts of this model, results of the original implementation are discussed. Then, recent improvements and extensions, like the inclusion of planetary migration and the resulting effects are discussed. It is shown that these improvement solve the ``timescale problem''. Finally, it is shown that by means of generating synthetic populations of (extrasolar) planets, core accretion models are able to reproduce in a statistically significant way the actually observed planetary population. ", "introduction": "Our current understanding of planet formation is based on several centuries of observations of the planets of our own Solar System, 12 years of extrasolar planets detection, and several decades of observations of young stellar systems. These studies have let to the general concept that after the collapse of a dense gas cloud, a protostar surrounded by a protoplanetary disk was formed. In this disk, solids started to coagulate from fine dust and grew further by mutual collision to form planetesimals (provided the bottleneck by bodies roughly one meter in size can be overcome), then protoplanets, and ultimately the actual planets. Some of the protoplanets managed to accrete a massive gaseous envelope onto their core. This is the very rough outline of the core accretion model. ", "conclusions": "The core accretion paradigm explains in an unified way the formation of giant and terrestrial planets, so that there is no need for a special mechanism for giant planets. Since the first core accretion models, significant improvement and extensions were made. Such improved and extended core accretion models can form giant planets well within observed disk lifetimes, so that there is no need for a faster formation mechanism. They have also reached a degree of maturity that allows quantitative tests with observations, of both the giant planets of our own Solar System, and, by means of population synthesis, of the extrasolar planet population. In the latter case, the whole population of detected planets can be used to constrain the models, which excludes model fine tuning for a specific case, and fully exploits the observational investment. As shown by statistical tests, extended core accretion models can reproduce many observed properties and correlations in the extrasolar planet population in a quantitative significant way with one synthetic population at one time. This means that accretion models can now be used to predict future observations, so that theory can feed back on the design of future instruments" }, "0710/0710.2047_arXiv.txt": { "abstract": "We review current state of neutron star cooling theory and discuss the prospects to constrain the equation of state, neutrino emission and superfluid properties of neutron star cores by comparing the cooling theory with observations of thermal radiation from isolated neutron stars. ", "introduction": "The equation of state (EOS) of superdense matter in neutron star cores is still a mystery. It is not clear if it is soft, moderate or stiff; if the matter contains nucleons/hyperons, or exotic components. In the absence of good practical theory of supranuclear matter the problem cannot be solved on purely theoretical basis, but it can be solved by comparing theoretical models with observations of neutron stars. The attempts to solve this long-standing problem by different methods are numerous (e.g., Refs.\\ \\cite{nsb1,lp07}). Here we discuss current results obtained from studies of cooling isolated neutron stars. The first papers on neutron star cooling appeared with an advent of X-ray astronomy, before the discovery of neutron stars. Their authors tried to prove that neutron stars cool not too fast and can be discovered as sources of thermal surface X-ray radiation. The first estimates of thermal emission from cooling neutron stars were most probably done by Stabler \\cite{stabler60} in 1960. Four years later Chiu \\cite{chiu64} made similar estimates and analyzed the possibility to discover neutron stars from their thermal emission. First, simplified calculations of neutron star cooling were done in 1964 and 1965 \\cite{morton64, cs64, bw65b}. The foundation of the strict cooling theory was laid in 1966 by Tsuruta and Cameron \\cite{tc66}, one year before the discovery of pulsars. We review the current state of the cooling theory. More details can be found in recent review papers \\cite{yp04,pgw06}. ", "conclusions": "The theory of cooling neutron stars of ages $10^2-10^6$ yr mostly tests the neutrino emission properties of the neutron star core. Its main results are as follows. (1) Neutrino emission in the outer core (i.e., in the core of a low-mass star) is a factor of 30--100 lower than the modified Urca emission in a nonsuperfluid star. (2) Neutrino emission in the inner core (of a massive star) is at least a factor of 30--100 higher than the modified Urca emission. It can be enhanced by direct Urca process in nucleon/hyperon inner core or by the presence of pion or kaon condensate, or quark matter. (3) The scenario with open direct Urca process predicts the existence (Fig.\\ \\ref{exotica}) of massive isolated neutron stars which are much colder than those observed now. In the scenario with pion condensate, the massive stars should be warmer (than those with open direct Urca) but colder than the observed ones. In the scenario with kaon condensate the massive stars should be even warmer but slightly colder than the observed sources. A discovery of cold cooling neutron stars would be crucial to constrain the level of enhanced neutrino emission in the inner core. (4) Observations of cooling neutron stars can be analyzed together with observations of SXTs in quiescent states. The data on SXTs indicate the existence of very cold neutron stars (first of all, SAX J1808.4--3658) which cool via direct Urca process, but the data and interpretation require additional confirmation. (5) A transition from slow neutrino emission in the outer core to enhanced emission in the inner core has to be smooth. Current observations of cooling neutron stars and SXTs do not constrain the parameters of this transition. A firm measurement of masses of cooling or accreting stars would help to impose such constraints. (6) New observations and reliable practical theories of dense matter are vitally important to tune the cooling theory as an instrument for exploring physical properties of neutron star interiors and neutron star parameters. The tuning will imply a careful analysis of many cooling regulators. \\begin{theacknowledgments} This work was partially supported by the Russian Foundation for Basic Research (grants 05-02-16245 and 05-02-22003), by FASI-Rosnauka (grant NSh 9879.2006.2), and by the Joint Institute for Nuclear Astrophysics (grant NSF PHY 0216783). \\end{theacknowledgments}" }, "0710/0710.5498_arXiv.txt": { "abstract": "We have observed nearly 200~FGK stars at 24~and 70~microns with the {\\em Spitzer} Space Telescope. We identify excess infrared emission, including a number of cases where the observed flux is more than 10~times brighter than the predicted photospheric flux, and interpret these signatures as evidence of debris disks in those systems. We combine this sample of FGK stars with similar published results to produce a sample of more than 350~main sequence AFGKM stars. The incidence of debris disks is 4.2$^{+2.0}_{-1.1}$\\% at 24~microns for a sample of 213~Sun-like (FG) stars and 16.4$^{+2.8}_{-2.9}$\\% at 70~microns for 225~Sun-like (FG) stars. We find that the excess rates for A, F, G, and K stars are statistically indistinguishable, but with a suggestion of decreasing excess rate toward the later spectral types; this may be an age effect. The lack of strong trend among FGK stars of comparable ages is surprising, given the factor of~50 change in stellar luminosity across this spectral range. We also find % that the incidence of debris disks declines very slowly beyond ages of 1~billion years. ", "introduction": "Planetary system formation must be studied through various indirect means due to the relative faintness of distant planets and the long timescales over which planetary system evolution takes place. One powerful technique has advanced substantially in the era of the {\\em Spitzer} Space Telescope: investigations of dusty debris disks around mature, main sequence stars. Debris disks arise from populations of planetesimals that remain from the era of planet formation; the analogs in our Solar System are the asteroid belt and the Kuiper Belt. Any system that possesses a debris disk necessarily has progressed toward forming a planetary system to some degree. The many small bodies that inhabit a debris disk can, on occasion, collide, producing a shower of fragments that grind each other down to dust particles. These dust grains can be heated by the central star to temperatures $\\sim$100~K, where they can be detected at wavelengths of 10--100~microns. Data from the IRAS and ISO satellites were used to identify and characterize debris disks \\citep[e.g.,][]{aumann84,decin00,spangler01,habing01,decin}. The Multiband Imaging Photometer for {\\em Spitzer} (MIPS; \\citet{mips}) offers substantially improved sensitivities at 24~and 70~microns and therefore can be used to advance the study of debris disks and measure the fraction of stars that possess colliding swarms of remnant planetesimals. A number of these surveys have been carried out using MIPS. \\citet{astars} and \\citet{astars2} observed hundreds of A~stars and found that the number of A~stars showing thermal infrared excess suggestive of collisionally-produced dust decreases as a function of stellar age, from 50\\% or more at ages just past the gas dissipation age of 10~Myr to 30\\% at 500~Myr. The excess rates are lower for older and lower mass stars. \\citet{bryden} found that the excess rate is around 15\\% for stars of mass and age similar to our Sun (for a sample of 69~stars). \\citet{gautier} did not detect any 24~or 70~micron excesses suggestive of debris disks in a sample of 60~field (old) M~stars (with only 13~strongly detected at 70~microns). Finally, multiplicity appears to play a significant role in modulating debris disks, as \\citet{binaries} found that the debris disk rate for A~and F~binaries was higher than that for single stars of similar spectral types. We present here a survey for excesses around almost 200~F, G, and K~stars, with ages and masses similar to that of our Sun. (Some of these data were published in \\citet{bryden}.) We characterize the excesses that we find and present a few systems of particular interest. We create a larger sample of FGK~stars by adding 75~stars from a similar survey, and derive excess rates as a function of spectral type and as a function of age. Finally, we discuss the implications of our results for planetary system formation. ", "conclusions": "\\subsection{Metallicity and excesses} Four stars in our FGK sample do not have published metallicities. For the 189~stars with known metallicities in Table~\\ref{targetinfo}, the mean metallicity is -0.08$\\pm$0.22, with a median of~-0.06. The mean metallicity of stars with excesses is -0.11$\\pm$0.19 (median is~-0.09). The mean metallicity of stars with no excesses is -0.08$\\pm$0.22 (median is~-0.05). There is no difference between the metallicity of the population of stars with excesses and the population of stars with no excesses. This lack of correlation has been discussed previously (e.g., B06, \\citet{chastpf}). \\subsection{Excess rates across spectral type \\label{spectral}} Many parameters affecting infrared excesses change with spectral type, including the importance of grain loss mechanisms (winds, Poynting-Robertson drag, photon pressure); stellar luminosity; and the locations of key temperatures (e.g., the ice line) in the systems. Significant surveys for debris disks across spectral types A--M~have now been published, and we use those data to look for systematic trends across spectral types. It is well known that excess rates decrease with stellar age \\citep{habing01,spangler01,astars,astars2,siegler}. This dependence must be avoided in testing for changes with spectral type. We take the oldest ($\\ge$600~Myr) A~stars from the \\citet{astars2} sample as our representative sample from that spectral type. For our F, G, and K~samples we take the union of the data presented here and the data in \\citet{chastpf}. The targets and data reduction presented in \\citet{chastpf} are quite similar to the selections and techniques we have employed here, which allows us to merge the two samples relatively seamlessly. We calculate the excess ratios for these F, G, and K samples and take the M~stars excess rates from \\citet{gautier}. This compilation is presented in Table~\\ref{excesssum} and Figure~\\ref{spexcess}. Within the error bars, the excess rates for the A, F, G, and K~subsamples are essentially indistinguishable. However, there is a suggestion of a trend of decreasing 70~\\micron\\ excess rates with later spectral types (Figure~\\ref{spexcess}). We note, however, that the mean age for the populations increases with later spectral types. It is possible that we are instead detecting a time-related effect, although the decay timescales identified by \\citet{astars}, \\citet{astars2}, and \\citet{siegler} of hundreds of millions of years should long since have diminished all disks at ages of billions of years. We discuss this possibility in Section~\\ref{ages}. \\citet{chastpf} remarked that the excess rates for K~stars appeared to be lower than that for F and G stars, finding zero excesses among 23~stars later than K2 (and zero excesses among a larger combined sample of 61~K1--M6~stars). We include the \\citet{chastpf} data in our analysis here, and find that, formally, the excess rate for K~stars is not significantly different than the excess rates for earlier (F and G) stars. As \\citet{chastpf} note, and we confirm, none of the 6~K~stars with excesses in our larger sample are later than K2. Part of the motivation for assembling the F~stars program (PID~30211) was as a control sample for the binary star program presented in \\citet{binaries}. In that study, 69~A3--F8 binary star systems were found, overall, to have relatively high excess rates: 9\\% at 24~microns and 40\\% at 70~microns. It is clear from our results here (see Figure~\\ref{spexcess}) that the (single) F~stars excess rate is equal to or lower than the (single) A~stars excess rate. The binaries excess rates remain significantly high compared to the control sample of A~and F~stars. Figure~\\ref{fd} shows that, in general, there is no trend of dust distance as a function of stellar effective temperature (spectral type). This may suggest that the processes that drive planetesimal formation do not depend strongly on a single critical temperature, as would be the case in the ``ice line'' model, where protoplanetary disk surface densities increase across certain temperature boundaries. However, our method for calculating dust distances may be too crude to see this effect. We note that all of the (minimum) dust distances given in Table~\\ref{excesstable} would fall within the planetary realm of our Solar System ($<$30~AU). We are not observing disks that are far outside of the potential planetary realm of these systems. We stated above that there is no particular trend for either fractional luminosity or dust distance with spectral type, but there is an important caveat: no disks with large dust distances or relatively small fractional luminosities were identified among the latest stars in our sample. This dearth may simply allude to the fact that later stars are cooler. Dust at 25~AU around a K1 (5000~K) star would have a temperature around 50~K; this dust would have its peak emission near 70~microns, and cooler (more distant) dust would have its peak emission longward. MIPS 70~micron observations of such a dust population, or a cooler one, would not readily show the presence of this excess. The lack of distant disks around K stars may therefore be an observational bias. Similarly, low fractional luminosity disks would be more difficult to detect around K~stars than around earlier stars, so the lack of low fractional luminosity disks for later stars may also be due to observational bias. These observational biases may be corrected with sufficiently sensitive measurements at $\\sim$100~microns (e.g., with Herschel). To further address the question of whether there is any detectable trend of excess rate as a function of spectral type using existing published data, we compared the incidence of 70~$\\mu$m excesses across these 5~spectral types (A, F, G, K, M), a total sample size of more than 350~stars. To avoid uncontrolled selection effects, we confined the comparison to stars that would have been detected at 70~microns at a level of at least 2:1 on the photosphere. We then applied a number of tests. First, we computed weighted average values of R70 (observed flux over predicted flux) as in \\citet{gautier}. We find a value of $\\sim$5 for the A stars, but we discount this large average because of the small size of the sample (27~stars). The values for the F, G, K, and M stars are 1.16, 1.23, 1.06, and 1.025, respectively. Errors are difficult to estimate because the excesses are not normally distributed. We also computed straight averages of R70 for these same samples. Here, we calculate 4, 2.6, 1.8, 1.4, and 1.1 for the A, F, G, K, and M stars, respectively. Again, the values need to be interpreted with caution because error estimation is difficult. To circumvent the difficulties in determining errors, we binned the excesses into intervals of 0.5~in excess ratio and used the K-S test to determine if the resulting distributions were likely to have been drawn from the identical parent distribution. For each stellar type, we tested the relevant distribution against the distribution for all the stars, excluding the contribution of the spectral type in question. The result was a set of probabilities of 0.03, 0.8, 0.3, 0.3, and 0.01 that the types A, F, G, K, and M, respectively, were drawn from the same parent distribution as the other types. (For this test, a claim of a significant difference requires a probability of 0.05 or less that the samples are from the same distribution. Values of~0.3 imply that the the G~and K~stars are 1$\\sigma$~different from their respective control samples.) From this suite of tests, we conclude that the incidence of excesses is different for old A stars and for M stars from that of the rest of our sample. We further deduce that there is a possibility of a difference appearing in the K~stars from their lower average excess ratio. The distributions for F and G stars appear to be indistinguishable with our data. We employ this conclusion in the creation of an FG ``supersample'' (Section~\\ref{supersample}). The higher excesses indicated for the old A stars could be an age effect, since the sample is by necessity significantly younger than the later types (which we selected in general to be $>$ 1 Gyr in age). In fact, \\citet{gorlova} and \\citet{siegler} compare 24$\\mu$m excesses from young A and solar-like stars and find that the incidence is quite similar at a given age. Age effects would presumably cause an apparent decrease of excess incidence with later type among the F, G, and K stars because F stars will evolve off the main sequence quickly enough to bias our sample toward younger objects (Section~\\ref{ages}). In the end, this discussion may still be suffering from a relatively small number of K~stars sampled. An ongoing {\\em Spitzer} program (PID~30490) to survey nearby stars that were not observed in other programs --- a sample that includes $\\sim$400~K~stars --- should help unravel these statistics. Nonetheless, given our result, it is unlikely that future surveys will find a strong trend among F, G, and K stars of comparable ages --- a range of spectral types that spans more than a factor of~50 in stellar luminosity. This behavior is counter to our expectations and is a challenge to models of debris disk evolution. \\subsection{Effects of age \\label{ages}} The lack of strong dependence on spectral type lets us combine data on various types to study the evolution with stellar age. Figure~\\ref{fgkage} shows individual R24 and $\\chi_{70}$ determinations for the stars in our FGK~sample whose ages are known, as a function of system age. There is no correlation apparent for these individual sources, so we look to binned data in a larger combined sample for evidence of trends. We once again take the union of the data presented here and that presented in \\citet{chastpf}, but this time we include only spectral types F0--K5 from the \\citet{chastpf} sample (that is, we exclude the latest K~stars). This is because the data we present in this paper covers the range F0--K5, and we want the best match to our combined sample. There are 10~stars in the F0--K5 TPF/SIM subsample that have no known ages, and 10~additional stars with ages less than 1~billion years. Of these~20, 2~systems have excesses (10\\%). Since this excess ratio is not significantly different from that of the overall FGK sample, and since the PID~30211 F~stars sample is controlled for age but the PID~41 sample is not specifically controlled for age, we make no attempt to correct the larger sample for age. Figure~\\ref{ageexcess} shows excess rate as a function of age for this combined sample. To zeroth order, there is no trend as a function of age: a constant excess rate of $\\sim$20\\% adequately fits the data, being consistent at 1$\\sigma$ with the 10~Gyr data point and at 1.5$\\sigma$ with the 8~Gyr data point. Furthermore, the 10~Gyr bin has only 7~targets in it, and the 8~Gyr bin has only 33~targets in it (still a relatively small number). On the other hand, we note that the data shown in Figure~\\ref{ageexcess} is suggestive of an excess rate that decreases with time through at least 8~Gyr. In this scenario, the 10~Gyr bin would be highly anomalous, although we note that this bin is clearly affected by small number statistics (2~excesses out of 7~stars). In other words, there may be a real trend and a real evolution of planetary systems and debris disks even on the billion-year timescale. However, this may again be a manifestation of the (potential) observational bias shown in Figure~\\ref{fd}, as follows. The fraction of stars in a given age bin that are K~stars increases for the later age bins. If K~stars truly have fewer excesses (or fewer detectable excesses, according to observational biases) than other spectral types, Figures~\\ref{ageexcess} and~\\ref{agesfd} might indeed be showing a real decrease with increasing age, but caused not by a long-timescale evolution of planetary systems but by the increasing dominance of excess-deficient K~stars at the oldest ages. The data present in our larger sample cannot distinguish between the competing possibilities of K~stars preferentially lacking (detectable) disks, or of old stars increasingly lacking disks. It also remains to be seen whether the high excess rate for the oldest bin in Figure~\\ref{ageexcess} is anything more than a small number statistics anomaly. The overall high rate of excess incidence in our samples (17\\%) indicates that the 400~Myr decay timescale the drives the evolution of A~star debris disks \\citep{astars2} cannot drive the evolution of debris disks in our $>$1~Gyr Sun-like sample. Instead, Sun-like stars appear to have a relatively constant incidence of 15\\%--20\\% that is not strongly dependent on age, but may be weakly dependent on age through a very long timescale decrease. \\subsection{Debris disks around Sun-like stars \\label{supersample}} Because there is no difference in excess rate between F and G stars, we can combine our sample of F0--G9~stars into a single population of ``Sun-like'' stars. To this sample of 169~stars % we add the 56~F~and G~stars from \\citet{chastpf} to create an even larger sample of 225~Sun-like stars (213~stars at 24~microns). We refer to this merged sample of 213~and 225~stars as the ``Sun-like supersample.'' The excess rates for this supersample are 4.2$^{+2.0}_{-1.1}$\\% at 24~microns % and 16.4$^{+2.8}_{-2.9}$\\% at 70~microns % (Table~\\ref{excesssum}). With this large supersample, we can now state the debris disk incidence rate for Sun-like stars with quite good confidence (small error bars). \\subsection{Implications for planetary system formation} The majority of the debris disk systems that we present here have excesses at 70~microns only, suggesting temperatures $\\lesssim$100~K and therefore an inner edge to the disks. These dusty debris disks are likely produced by collisions within a swarm of planetesimals akin to the asteroid belt or Kuiper Belt in our Solar System. We interpret the cool temperatures we derive for the dust in these disks as evidence of inner disk holes where the surface density of dust is much smaller than in the planetesimal ring, and potentially zero. This architecture is strongly reminiscent of our Solar System, where planets sculpt the edges of the planetesimal and dust belts. It may be that many of the systems we discuss here similarly have planets sculpting their dust distributions. In several cases, the known planets may indeed be the ones sculpting the inner edges of the disks (Tables~\\ref{colortemps} and~\\ref{excesstable}). Additionally, Lawler et al.\\ (in prep.) have found, using IRS spectra, that the incidence of detectable levels of warm dust may be higher than the 4.2\\% we find here, indicating that dust in a region analogous to our Solar System's asteroid belt may also be somewhat common. \\citet{bryden07} show that the properties of disks around planet-bearing stars are unlikely to be similar to the properties of disks around stars without known planets. Briefly, the detection rates between these two populations are similar, but the planet-bearing stars are generally farther away and in more confused regions of the sky. They conclude that disks around planet-bearing stars are dustier (i.e., more massive), and probably more common, than disks around stars without known planets. Additionally, \\citet{binaries} recently showed that the excess rate for binary A-F~stars is 9\\% and 40\\% at 24~and 70~\\micron, respectively, significantly higher than our results for FGK stars and for Sun-like stars. We see in Figure~\\ref{spexcess} that these excess rates are relatively high, compared to the large sample of single stars we present here. Combining these two studies, we find strong evidence that the presence of additional massive bodies (whether stellar or planetary companions) in stellar systems appears to promote higher excess rates. This effect may simply be dynamical --- more massive bodies means more stirring, more collisions, and more dust --- but it is not clear that such a process could remain effective for billions of years. Further work is necessary to explain these results. Figure~\\ref{spexcess} shows that the transition from the high excess rate around A~stars to the more modest excess rate around Sun-like stars is gradual. This gradual decay is likely an age effect (consider the mean ages of the samples given in Figure~\\ref{spexcess}). Our data show that excess rates for FGK stars decline slowly with stellar age, and it is not clear whether we are detecting a billion-year tail of debris disk evolution, a dependence on spectral type, and/or an observational bias. Young ($<$1~Gyr) debris disks decay on 100--400~million year timescales \\citep{astars,astars2,gorlova,siegler}. The presence of excesses around 16\\% of Sun-like stars at billion year ages indicates that these old debris disks must be driven by a different evolutionary process than debris disks around those younger stars. One interpretation, using our Solar System as an analogy, is that the younger systems are still active in a Late Heavy Bombardment kind of dynamical upheaval. After a billion years, such large scale processes likely have ceased in all but the most unusual systems. Debris disks are then produced from collisions within remaining planetesimal belts (e.g., our asteroid belt) that have been dynamically excited by the previous eon's dynamical stirring. However, there is no trend of fractional luminosity with age (Figure~\\ref{agesfd}), although there is a lack of high fractional luminosity disks at old ages. The presence of a debris disk indicates that planetary system formation progressed at least to the planetesimal stage in a given system. There is no apparent dependence on spectral type across Sun-like stars for fractional luminosity (i.e., dust mass), dust distance, or even excess rate. This indicates that the processes that give rise to debris disk --- planetesimal formation, dynamical stirring that produces collisions --- must equally be insensitive to stellar parameters (temperature, mass, luminosity). This may argue that planetary system formation is quite robust --- able to occur in many different conditions. While none of the debris disks we observed are very similar to our own Solar System, there can be no question that the process of planetary system formation is quite common." }, "0710/0710.4048.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory % \\abstract % context heading (optional) % {} % % % aims heading (mandatory) % {For Seyfert galaxies, the AGN unification model provides a simple and well established explanation of the Type~1/Type~2 dichotomy through orientation based effects. The generalization of this unification model to the higher luminosity AGNs that are the quasars remains a key question. The recent detection of Type~2 Radio-Quiet quasars seems to support such an extension. We propose to further test this scenario.} % {Type~1 but that they are seen at smaller inclination preventing us to see the broad % emission lines present in the spectrum of Type~1 AGN.} % % % methods heading (mandatory) % pola + deconvo we search ... {On the basis of a compilation of quasar host galaxy position angles consisting of previously published data and of new measurements performed using HST Archive images, we investigate the possible existence of a correlation between the linear polarization position angle and the host galaxy/extended emission position angle of quasars.} %% % % results heading (mandatory) {We find that the orientation of the rest-frame UV/blue extended emission is correlated to the direction of the quasar polarization. For Type~1 quasars, the polarization is aligned with the extended UV/blue emission while these two quantities are perpendicular in Type~2 objects. This result is independent of the quasar radio-loudness. We interpret this (anti-)alignment effect in terms of scattering in a two-component polar+equatorial model which applies to both Type~1 and Type~2 objects. Moreover the orientation of the polarization --and then of the UV/blue scattered light-- does not appear correlated to the major axis of the stellar component of the host galaxy measured from near-IR images.} % % conclusions heading (optional), leave it empty if necessary {} % These observations can be simply explained in the framework of an % extension to quasars of the unification model of AGN where the polarization is produced in two scattering regions. %% ", "introduction": "The study of quasars shows that we can classify them among various categories. Radio-Loud quasars (RLQ) are distinguished from the Radio-Quiet quasars (RQQ) according to their radio power (Kellerman et al. \\cite{ke89}), and the Type~1/Type~2 objects from the presence or absence of broad emission lines in their spectrum (Lawrence \\cite{law87}). One can then wonder whether there exists a common link between all these objects, i.e. are the physical processes at the origin of all quasars the same? An interesting way to tackle this question consists in the use of the linear optical polarization. Polarimetry, in combination with spectroscopy, has led to major advances in the development of a unified scheme for AGN (see Antonucci \\cite{anto93} for a review). The discovery of hidden broad emission lines in the polarized spectrum of the Type~2 Seyfert galaxy NGC1068 led to consider these objects as intrinsically identical to Type 1 Seyfert, the edge-on orientation of a dusty torus blocking the direct view of the central engine and the broad emission line region (Antonucci \\& Miller \\cite{anto85}). The question we investigate in this paper relates to the possible existence of a correlation between the optical polarization position angle $\\theta_{Pola}$\\footnote{The polarization position angle $\\theta_{Pola}$ is defined as the position angle of the maximal elongation of the electric vector in the plane of polarization, measured in degrees East of North.} and the orientation of the host galaxy/extended emission $PA_{host}$\\footnote{The orientation of the host galaxy $PA_{host}$ is characterized by the position angle of its major axis projected onto the plane of the sky, measured in degrees East of North.} in the case of RLQs and RQQs. The relation between quasars and their host galaxies may play a fundamental role in our understanding of the AGN phenomenon and in determining the importance of the feedback of AGN on their hosts. Such a correlation has already been studied in the case of the less powerful AGN that are the Seyfert galaxies. Thompson \\& Martin (\\cite{toma88}) found a tendency for Seyfert 1 to have their polarization angle aligned with the major axis of their host galaxy, an observation that they interpreted as due to dichroic extinction by aligned dust grains in the Seyfert host galaxy. In the case of the more powerful AGN that are quasars, Berriman et al. (\\cite{beri90}) investigated this question. They determined by hand the $PA_{host}$ of 24 PG quasars from ground based images and computed the acute angle $\\Delta \\theta$ between $\\theta_{Pola}$ and $PA_{host}$. They observed that while more objects seem to appear at small values ($\\Delta \\theta \\leq 45\\degr$), this effect was only marginally statistically significant. Investigating this problem with ground based data for Type~1 quasars is hampered by the inability of separating the faint host galaxy/extended emission\\footnote{In the following, we use indifferently the term host galaxy or extended emission to refer to all the extended emission around the central source including stars, ionized gas or scattered light.} from the blinding light of the quasar nucleus. Because of its high angular resolution and stable Point Spread Function (PSF), the Hubble Space Telescope (HST) permits to properly remove the contribution of the powerful quasar nucleus thus allowing the investigation of the host galaxy parameters. A large number of quasar host observing programs were carried out with the HST leading to a statistically useful sample of quasar host galaxy parameters (e.g. Bahcall et al. \\cite{ba97}; Dunlop et al. \\cite{du03} and many more see Sects.~\\ref{publidata} and \\ref{newosdat}). Our aim is to investigate the $\\theta_{Pola}/PA_{host}$ relation on the basis of high resolution HST quasar images. We will use either position angles given in the literature or $PA_{host}$ we measured ourselves from HST Archive observations (hereafter called ``\\emph{new $PA_{host}$ data}\"). The layout of this paper is as follows. In Sect.~\\ref{publidata} we introduce the samples of quasars used in this study which possess a $PA_{host}$ given in the literature. In Sect.~\\ref{tres}, we outline the samples with good imaging data but for which no $PA_{host}$ were published, and we summarize our data analysis process and the approach followed to model the HST Archive images and to derive the host galaxy parameters. In Sect.\\ref{rapola} we briefly describe the polarization and radio data. Then in Sect.~\\ref{statos} we present the statistical analysis of the sample and the results obtained. In Sect.~\\ref{discu} we discuss the results and compare them to former studies. Finally, our conclusions are summarized in Sect.~\\ref{conclu}. %__________________________________________________________________ ", "conclusions": "\\label{conclu} Using host galaxy position angles ($PA_{host}$) determined from high resolution optical/near-IR images of quasars and data from the literature, we investigate the possible existence of a correlation between the host morphology and the polarization direction in the case of Radio-Quiet and Radio-Loud quasars. We can summarize our results as follows : \\begin{enumerate} \\item We find an alignment between the direction of the linear polarization and the rest-frame UV/blue major axis of the host galaxy of Type~1 quasars. In the case of Type~2 objects, it is well established that the extended UV/blue light is correlated to the observed polarization, as these blue regions are thought to be dominated by scattering. Our results suggest that such an extended UV/blue scattering region is also present in Type~1 quasars. \\item We do not find such an alignment effect with the near-IR host morphology. This suggests that the morphology of the extended UV/blue emission is not related to the morphology of the stellar component of the host galaxy which dominates in the near-IR. \\item We observe the same $PA_{host} - \\theta_{Pola}$ behavior for either Radio-Loud or Radio-Quiet objects. This observation supports the idea that the UV/blue continuum is not entirely due to star formation processes triggered by the radio-jet. \\item The observed correlation fits a unification model where the Type~1/Type~2 dichotomy is essentially determined by orientation effects assuming the two component scattering model of Smith et al. (\\cite{sm04,sm05}). Indeed, depending on the viewing angle to the quasar, the polarization would predominantly arise from either the equatorial or the polar scattering region giving rise to the observed behavior. \\end{enumerate} In order to strengthen the conclusions and to further investigate the correlations presented in this paper, new observations of quasars in the rest-frame UV/blue domain are needed, especially of Radio-Quiet objects for which few observations at $\\lambda_{rest} \\le 5000$ \\AA~are available. The detection of the extended blue/UV continuum region and the measurement of its polarization would help to test our interpretation." }, "0710/0710.5721_arXiv.txt": { "abstract": "The equation of state for radiation is derived in a canonical formulation of the electromagnetic field. This allows one to include correction terms expected from canonical quantum gravity and to infer implications to the universe evolution in radiation dominated epochs. Corrections implied by quantum geometry can be interpreted in physically appealing ways, relating to the conformal invariance of the classical equations. ", "introduction": "\\label{sec:INTRODUCTION} In theoretical cosmology, many insights can already be gained from spatially isotropic Friedmann--Robertson--Walker models \\begin{equation} \\md s^2= -\\md\\tau^2+a(\\tau)^2\\left(\\frac{\\md r^2}{1-kr^2}+r^2(\\md\\vartheta^2+ \\sin^2\\vartheta\\md\\varphi^2)\\right) \\end{equation} with $k=0$ or $\\pm 1$. The matter content in such a highly symmetric space-time can only be of the form of a perfect fluid with stress-energy tensor $T_{ab}=\\rho u_au_b+P (g_{ab}+u_au_b)$ where $\\rho$ is the energy density of the fluid, $P$ its pressure and $u^a$ the 4-velocity vector field of isotropic co-moving observers. Once an equation of state $P=P(\\rho)$ is specified to characterize the matter ingredients, the continuity equation $\\dot{\\rho}+3H(\\rho+P)=0$ with the Hubble parameter $H=\\dot{a}/a$ allows one to determine the behavior of $\\rho(a)$ in which energy density changes during the expansion or contraction of the universe. This function, in turn, enters the Friedmann equation $H^2+k/a^2=8\\pi G\\rho/3$ and allows one to derive solutions for $a(\\tau)$. In general, one would expect the equation of state $P=P(\\rho)$ to be non-linear which would make an explicit solution of the continuity and Friedmann equations difficult. It is thus quite fortunate that in many cases linear equations of state $P=w\\rho$ with $w$ constant are sufficient to describe the main matter contributions encountered in cosmology at least phenomenologically. The influence of compact objects on cosmological scales is, for instance, described well by the simple dust equation of state $P(\\rho)=0$. Relativistic matter, mainly electromagnetic radiation, satisfies the linear equation of state $P=\\frac{1}{3}\\rho$. The latter example is an exact equation describing the Maxwell field, rather than an approximation for large scale cosmology. It is thus, at first sight, rather surprising that the dynamics of electromagnetic waves in a universe can be summarized in such a simple equation of state irrespective of details of the field configuration. The result follows in the standard way from the trace-freedom of the electromagnetic stress-energy tensor and is thus related to the conformal symmetry of Maxwell's equations. That the availability of such a simple equation of state is very special for a matter field can be seen by taking the example of a scalar field $\\phi$ with potential $V(\\phi)$. In this case, we have an energy density $\\rho=\\frac{1}{2}\\dot{\\phi}^2+V(\\phi)$ and pressure $P=\\frac{1}{2}\\dot{\\phi}^2-V(\\phi)$. Unless the scalar is free and massless, $V(\\phi)=0$ for which we have a stiff fluid $P=\\rho$, there is no simple relation between pressure and energy density independently of a specific solution. Any conformal symmetry such as that of elecromagnetism might be broken by quantum effects especially when quantum gravity with its new scale provided by the Planck length is taken into account. The coupling of the electromagnetic field to geometry will then change, and exact conformal symmetries can easily be violated. Accordingly, one expects corrections from quantum gravity to the radiation equation of state and corresponding effects in the universe evolution during radiation dominated epochs. In loop quantum cosmology \\cite{LivRev} equations of state of matter fields are in general modified by perturbative corrections at large scales and non-perturbative ones on small scales \\cite{InvScale}. This has mainly been studied so far for a scalar field for which quantum modifications can be so strong that negative pressure results independently of the chosen potential \\cite{Inflation}. The main reason is the fact that the isotropic scalar field Hamiltonian $H_{\\phi}=\\frac{1}{2}a^{-3}p_{\\phi}^2+ a^3V(\\phi)$, where $p_{\\phi}$ is the momentum of $\\phi$, contains an inverse power of the scale factor $a$. For quantum gravity, this factor has to be quantized, too. Using the methods of \\cite{QSDV}, it turns out that inverse powers receive strong loop quantum corrections at small length scales \\cite{InvScale}. Accordingly, such modifications play a role for effective equations describing the universe after the big bang (or even during the quantum transition through the big bang singularity). During later stages, modifications are expected to decrease in size, but they might still be relevant due to sometimes tight constraints on evolution parameters. An extension to the usual matter ingredients of cosmology with linear equations of state is, however, difficult since the modification is based on quantizations of the fundamental field Hamiltonians. Equations of state are obtained from fundamental Hamiltonians after an analysis of the matter field equations, which can be difficult in general especially when quantum effects are taken into account. The only exception is the dust case since it implies a constant Hamiltonian (the total mass of dust) which is straightforwardly quantized without any corrections. Thus, although the dust energy density is proportional to $a^{-3}$ and metric dependent in a way which involves the inverse, it does not receive any modification since the Hamiltonian, i.e.\\ total energy $a^3\\rho$, is the essential object to be quantized. For radiation with $\\rho\\propto a^{-4}$ the expectation is not clear since the total energy does behave like an inverse power of $a$, but this follows only after an indirect analysis of the field dynamics. It is not the solution $\\rho(a)\\propto a^{-4}$ of the continuity equation which is quantized but the original field Hamiltonian from which the equation of state has to be derived first. One thus has to go back to the fundamental Maxwell Hamiltonian, derive energy density and pressure and see how quantum effects change the equation of state. If this is completed, one may attempt to solve the continuity equation to obtain corrections to $\\rho(a)$. We will derive such corrections in this article, using the canonical quantization given by loop quantum gravity \\cite{Rov,ALRev,ThomasRev}. Candidates for Hamiltonian operators of the Maxwell field have been proposed \\cite{QSDV} which show several sources of correction terms. To derive corrections to the equation of state, however, we need to perform the usual calculation in a Hamiltonian formulation. Thus, we first present the canonical formulation for the free classical Maxwell field to rederive the standard result for the equation of state parameter $w$ without reference to an action or the stress-energy tensor. Appropriate modifications to the matter Hamiltonian $H_{M}$ are then made to derive possible loop quantum gravity corrections to the equation of state $w$. We will show that one case of corrections results again in a linear equation of state, albeit in a corrected way which depends on the basic discreteness scale of quantum gravity. In this case we are able to express, as in the classical case, the full field dynamics in terms of a simple modified $w$, and to solve explicitly for $\\rho(a)$. Our derivation takes into account inhomogeneous field configurations and presents the first modified equation of state obtained for a realistic matter source in loop quantum gravity. ", "conclusions": "\\label{sec:DISCUSSIONS} We have derived here the equation of state of the Maxwell field in a canonical form, including corrections expected from loop quantum gravity. In the canonical derivation, the reason for a linear equation of state, which is trace-freedom in the Lagrangean derivation, is the fact that the same metric dependent factor $q_{ab}/\\sqrt{q}$ multiplies both terms in the Hamiltonian. The Maxwell Hamiltonian is thus simply rescaled if the metric is conformally transformed, which explains the conformal invariance of Maxwell's equations. This is special for the Maxwell field and different from, e.g., a scalar field with a non-vanishing potential. The same fact allows one to quantize the Hamiltonian in a way which affects both the electric and magnetic term in the same way, at least as far as the metric dependence is concerned. One then obtains a single correction function $\\alpha=\\beta$ which only corrects the metric dependence of the total scale of the Hamiltonian. In this sense, conformal invariance is preserved even after quantization. (But this would not be the case if a quantization is used which results in $\\alpha\\not=\\beta$.) This preservation of the form of the Hamiltonian explains why we are still able to derive an equation of state independently of the specific field dynamics and that it remains linear. However, the classical value $w=\\frac{1}{3}$ is corrected due to quantum effects in the space-time structure. This modification is also understandable from a Lagrangean perspective, together with basic information from the loop quantization. Employing trace freedom of the stress-energy tensor to derive the equation of state, we have to use the inverse metric in $g^{ab}T_{ab}$. But from loop quantum gravity we know that, when quantized, not all components of the inverse metric agree with inverse operators of the quantization. For the scale factor of an isotropic metric, for instance, we have $\\widehat{a^{-1}}\\not=\\mbox{``$\\hat{a}^{-1}$''}$ since the right hand side is not even defined \\cite{InvScale}. While the left hand side is defined through identities such as (\\ref{poissonbracketofvolume}), it satisfies $\\widehat{a^{-1}}\\hat{a}\\not=1$ and thus shows deviations from the classical expectation $a^{-1}a=1$ on small scales which were captured here in correction functions. As derived in detail, this implies scale dependent modifications to the equation of state parameter $w_{\\rm eff}$. The result can also be interpreted in more physical terms. The classical behavior $\\rho(a)\\propto a^{-4}$ can be understood as a combination of a dilution factor $a^{-3}$ and an additional redshift factor $a^{-1}$ for radiation in an expanding universe. As we have seen, this is corrected to $\\alpha(a)a^{-4}$ where $\\alpha(a)$ corrects the metric factor $q_{ab}/\\sqrt{q}\\sim a^{-1}\\delta_{ab}$. Since this is only a single inverse power of $a$ for an isotropic solution, we can interpret the result as saying that only redshift receives corrections due to quantum effects on electromagnetic propagation. The dilution factor due to expansion is unmodified, except that the background evolution $a(t)$ itself receives corrections. This agrees with the result for dust, which is only diluted and has an unmodified equation of state even after quantization \\footnote{But it disagrees with \\cite{Metamorph} both for dust and radiation, where a direct quantization of energy densities exclusively for isotropic fields was attempted.}. Unlike dust, for radiation one has to refer to the inhomogeneous field and its quantum Hamiltonian to derive a reliable equation of state, as presented here." }, "0710/0710.5517_arXiv.txt": { "abstract": "In a large region of the supersymmetry parameter space, the annihilation cross section for neutralino dark matter is strongly dependent on the relative velocity of the incoming particles. We explore the consequences of this velocity dependence in the context of indirect detection of dark matter from the galactic center. We find that the increase in the annihilation cross section at high velocities leads to a flattening of the halo density profile near the galactic center and an enhancement of the annihilation signal. ", "introduction": " ", "conclusions": "" }, "0710/0710.2930_arXiv.txt": { "abstract": "Hot Jupiters are new laboratories for the physics of giant planet atmospheres. Subject to unusual forcing conditions, the circulation regime on these planets may be unlike anything known in the Solar System. Characterizing the atmospheric circulation of hot Jupiters is necessary for reliable interpretation of the multifaceted data currently being collected on these planets. We discuss several fundamental concepts of atmospheric dynamics that are likely central to obtaining a solid understanding of these fascinating atmospheres. A particular effort is made to compare the various modeling approaches employed so far to address this challenging problem. ", "introduction": "An exploding body of observations promises to unveil the meteorology of giant planets around other stars. Because of their high temperatures, short orbital periods, and likelihood of transiting their stars, the hot Jupiters are yielding their secrets most easily, and we now have constraints on radii, composition, albedo, dayside temperature structure, and even day-night temperature distributions for a variety of planets \\citep[e.g.,][] {knutson-etal-2007b, cowan-etal-2007, harrington-etal-2006, harrington-etal-2007, charbonneau-etal-2002, charbonneau-etal-2005, charbonneau-etal-2007, deming-etal-2005, deming-etal-2006}. A knowledge of atmospheric dynamics is crucial for interpreting these observations. Understanding the atmospheric circulation of {\\it any} planet is a difficult task, and hot Jupiters are no exception. The difficulty results from the nonlinearity of fluid motion and from the complex interaction between radiation, fluid flow, and cloud microphysics. Turbulence, convection, atmospheric waves, vortices, and jet streams can all interact in a complex manner across a range of temporal and spatial scales. Furthermore, radiative transfer and dynamics are coupled and cannot be understood in isolation. For example, the equator-to-pole temperature contrasts on terrestrial planets depend not only on the rate of latitudinal energy transport by the circulation but also on the way the atmospheric radiation field is affected by the advected temperature field. Cloud microphysics complicates the problem even more, because the large-scale radiative properties of clouds depend on the microphysics of the cloud particles (particle number density, shape, size distribution, and absorption/scattering properties). We thus have a non-linear, coupled radiation-hydrodynamics problem potentially involving interactions over 14 orders of magnitude, from the cloud-particle scale ($\\sim1\\,\\mu$m) to the global scale ($\\sim10^8\\,$m for hot Jupiters). It is worth emphasizing here the important difference between the complexity of planetary atmospheres and the relative simplicity of stars, exemplified by the main sequence in the HR diagram. While specifying a few global parameters---such as mass, composition, and age---is typically sufficient to understand the key observable properties of stars, this is generally not the case for planetary atmospheres. It is possible that a significant observable diversity exists even among a group of extrasolar planets which share similar global attributes. Understanding such complexity is a challenge for extrasolar planetary science. A proven method for dealing with the complexity of planetary atmospheres is the concept of a {\\it model hierarchy.} Even a hypothetical computer model that included all relevant processes and perfectly simulated an atmosphere would not, by itself, guarantee an {\\it understanding} of the atmospheric behavior any more than would the observations of the real atmosphere \\citep{held-2005}. This is because, in complex numerical simulations, it is often unclear how and why the simulation produces a specific behavior. To discern which processes cause which outcomes, it is important to compare models with a range of complexities (in which various processes are turned on or off) to build a hierarchical understanding. For example, east-west jet streams occur in the atmospheres of all the planets in our Solar System. The study of these jets involves a wide range of models, all of which have important lessons to teach. The most idealized are pure 2D models that investigate jet formation in the simplest possible context of horizontal, 2D turbulence interacting with planetary rotation \\citep[e.g.,][]{williams-1978, yoden-yamada-1993, cho-polvani-1996b, huang-robinson-1998, sukoriansky-etal-2007}. Despite the fact that these simplified models exclude vertical structure, thermodynamics, radiation, and clouds and provide only a crude parameterization of turbulent stirring, they produce jets with several similarities to those on the planets. Next are one-layer ``shallow-water-type'' models that allow the fluid thickness to vary, which introduces buoyancy waves, alters the vortex interaction lengths, and hence changes the details of jet formation \\citep{cho-polvani-1996a, cho-polvani-1996b, scott-polvani-2007, showman-2007}. 3D models with simplified forcing allow the investigation of jet vertical structure, the interaction of heat transport and jet formation, and the 3D stability of jets to various instabilities \\citep{cho-etal-2001, williams-1979, williams-2003a, schneider-2006, lian-showman-2007}. Such models suggest, for example, that Jupiter and Saturn's superrotating\\footnote{Superrotation is defined simply as a positive (eastward) longitudinally averaged wind velocity at the equator, so that the atmospheric gas rotates faster than the planet.} equatorial jets may require 3D dynamics. Finally are full 3D general-circulation models (GCMs) that include realistic radiative transfer, representations of clouds, and other effects necessary for detailed predictions and comparisons with observations of specific planets. The comparison of simple models with more complex models provides insights not easily obtainable from one type of model alone. This lesson applies equally to hot Jupiters: a hierarchy of models ranging from simple to complex will be necessary to build a robust understanding. Here we review our current understanding of atmospheric circulation on hot Jupiters. We first describe basic aspects of relevant theory from atmospheric dynamics; this is followed by a detailed comparison of the equations, forcing methods, and results obtained by the different groups attempting to model the atmospheric circulation on these fascinating objects. ", "conclusions": "The study of hot Jupiter atmospheres is maturing. While data of increasing quality are being collected, atmospheric models are also being refined to help build a robust, hierarchical understanding of these unusual atmospheres. Indeed, one of the main motivations for studying these atmospheres, which is also a main source of difficulty, is the unusual physical regime that characterizes them. Hence, hot Jupiters represent new laboratories for studying the complex physics of giant planet atmospheres. In this way, they offer the promise of extending the boundary of comparative planetology well beyond the solar-system planets." }, "0710/0710.4511_arXiv.txt": { "abstract": "We present a novel technique to phase-lock two lasers with controllable frequency difference. In our setup, one sideband of a current modulated Vertical-Cavity Surface-Emitting Laser (VCSEL) is phase locked to the master laser by injection seeding, while another sideband of the VCSEL is used to phase lock the slave laser. The slave laser is therefore locked in phase with the master laser, with a frequency difference tunable up to about 35 GHz. The sideband suppression rate of the slave laser is more than 30dB at 30 $\\mu$W seed power. The heterodyne spectrum between master and slave has a linewidth of less than 1 Hz. A narrow linewidth spectrum of coherent population trapping in rubidium is achieved using such beams. ", "introduction": " ", "conclusions": "" }, "0710/0710.3041_arXiv.txt": { "abstract": "A perturbative analysis is used to investigate the effect of rotation on the instability of a steady accretion shock (SASI) in a simple toy-model, in view of better understanding supernova explosions in which the collapsing core contains angular momentum. A cylindrical geometry is chosen for the sake of simplicity. Even when the centrifugal force is very small, rotation can have a strong effect on the non-axisymmetric modes of SASI by increasing the growth rate of the spiral modes rotating in the same direction as the steady flow. Counter-rotating spiral modes are significantly damped, while axisymmetric modes are hardly affected by rotation. The growth rates of spiral modes have a nearly linear dependence on the specific angular momentum of the flow. The fundamental one-armed spiral mode ($m=1$) is favoured for small rotation rates, whereas stronger rotation rates favour the mode $m=2$. A WKB analysis of higher harmonics indicates that the efficiency of the advective-acoustic cycles associated to spiral modes is strongly affected by rotation in the same manner as low frequency modes, whereas the purely acoustic cycles are stable. These results suggest that the linear phase of SASI in rotating core-collapse supernovae naturally selects a spiral mode rotating in the same direction of the flow, as observed in the 3D numerical simulations of Blondin \\& Mezzacappa (2007). This emphasizes the need for a 3D approach of rotating core-collapse, before conclusions on the explosion mechanisms and pulsar kicks can be drawn. ", "introduction": "} Despite extensive studies, the explosion mechanism of core-collapse supernovae is still elusive. According to the delayed explosion scenario, the shock is first stalled at a distance of a few hundred kilometers, and then revived after neutrinos diffuse out of the proto neutron star. Unfortunately, numerical simulations suggest that neutrino heating may not be efficient enough, at least in spherical symmetry \\citep{lieb05}. Recent studies have shown that the spherical stalled shock is unstable against non radial perturbations with a low degree $l=1,2$ even if the flow is convectively stable. This result was demonstrated using axisymmetric numerical simulations \\citep{blo03,blo06,sch06,sch08,ohn06} and linear stability analyses \\citep{gal05,fog07,yam07}. Some numerical simulations have shown that this hydrodynamical instability, often called SASI, may assist the revival of the shock and trigger a successful explosion, powered either by neutrino heating \\citep{mar07} or by acoustic waves \\citep{bur06}. Some observed properties of young neutron stars may also be the consequences of SASI, such as their distribution of velocities \\citep{sch04, sch06} or their spin \\citep{blo07a,blo07b}. Until now, most studies of SASI have assumed that the unperturbed flow is purely radial and not rotating. Since the angular momentum of massive stars is likely to be large \\citep{heg05}, it is desirable to understand how the properties of SASI are affected by rotation. In this Letter, the effect of rotation on the linear stage of SASI is investigated using a perturbative analysis in order to shed light on one of the surprising results observed by \\cite{blo07a} in their 3D numerical simulations: the development of SASI seems to systematically favour a spiral mode rotating in the same direction as the accretion flow. As a consequence of momentum conservation, this mode diminishes and may even reverse the angular momentum acquired by the proto-neutron star from the stationary flow. Incidentally, the present linear study does not address another surprising result of \\cite{blo07a}, that a spiral mode of SASI always dominate the axisymmetric mode even without rotation. Following an approach similar to \\citet{fog07} (hereafter FGSJ07), we first compute the eigenfrequencies by solving accurately a boundary value problem between the shock surface and the accretor surface; in a second step, we use the same WKB method as in FGSJ07 to measure the stability of purely acoustic and advective-acoustic cycles in this region. This approach is different from \\cite{lam07}, which is based on the approximate derivation of a dispersion relation. Rather than the complexity of describing the non-spherical shape of a shock deformed by rotation \\citep{yam05}, we have chosen, as a first step, to solve the much simpler problem of a cylindrical accretion shock. This flow is simple enough to allow for a complete coverage of the parameter space and a physical insight of the main effects of rotation on SASI. Once characterized, these effects can be transposed into the more complex geometry of a rotating stellar core. ", "conclusions": "} \\subsection{Instability mechanism} \\label{mechanism} As underlined in Sect.~3, the dynamical effect of the centrifugal force on the stationary flow is modest. We anticipated in Sect.~2 that the only linear effect of angular momentum is a Doppler shift of the eigenfrequency $\\omega'=\\omega-m\\Omega(r)$, where $\\Omega(r)$ is the local rotation frequency. This leaves the axisymmetric mode $m=0$ unaffected and explains the relative insensitivity of its growth rate with respect to the rotation rate, at least for moderate angular momentum. The strong effect of rotation on the growth rate of the spiral modes can thus be traced back to this Doppler shifted frequency. What is the mechanism of the instability ? As seen in the previous section, the destabilizing role of rotation does not seem related to the presence or absence of a corotation radius, thus discarding a Papaloizou-Pringle mechanism \\citep{gol85}. Two possibilities have been proposed for the mechanism of SASI without rotation; one is the advective-acoustic mechanism \\citep{fog00,fog01,fog02} and the other is the purely acoustic mechanism \\citep{blo06}. Up to now, there is no satisfactory direct argument for the mechanism of the modes with a long wavelength. FGSJ07 used a WKB approximation to prove that the instability of the modes with a short wavelength is due to an advective-acoustic mechanism and extrapolated this conclusion to the modes with a long wavelength, which are the most unstable. This method, recalled in Appendix~B, is based on the identification of acoustic waves and advected waves at a radius immediately below the shock surface, and the measurement of their coupling coefficients, above this radius due to the shock, and below this radius due to the flow gradients. These coupling processes are responsible for the existence of two cycles, namely a purely acoustic cycle characterized by an efficiency ${\\cal R}$, and an advective-acoustic cycle characterized by an efficiency ${\\cal Q}$. By using the same method, the present study does not address directly the instability mechanism of long wavelength modes. However, the WKB approximation enables us to describe, in a conclusive manner, the instability mechanism of short wavelength modes affected by rotation. First we checked that when the shock distance is increased ($r_{\\rm sh}=20r_*$), the overtones are also unstable and their growth rate is an oscillatory function of the frequency similar to Fig.~7 of FGSJ07. The effect of rotation on the advective-acoustic cycle is illustrated by Fig~3, for the spiral modes $m=\\pm1$ corresponding to the $10$-th overtone, as a function of the rotation rate. The cycle efficiency ${\\cal Q}$ is strongly amplified by rotation if $m>0$, while strongly damped if $m<0$. The stabilization of the counter-rotating spiral coincides with a marginally stable cycle ${\\cal Q}\\sim1$. The calculation of the amplification factor ${\\cal R}$ of perturbations during each purely acoustic cycle indicates its stability (${\\cal R}<1$). Contrary to the expectation of \\citet{lam07} (see next subsection), rotation clearly favours the spiral mode of the advective-acoustic cycle. This consequence of rotation established unambiguously for short wavelength perturbations is identical to the influence of rotation on the fundamental mode of SASI: we consider this a new hint that the advective-acoustic mechanism can be extrapolated to low frequencies. The detailed analysis of the consequences of the Doppler shifted frequency on the increase of the advective-acoustic efficiency ${\\cal Q}$ will be presented elsewhere (\\cite{yam08}, in preparation). \\subsection{Comments on the Results of Laming (2007)} \\label{comment} The effect of rotation on the growth rate of SASI, established in Sect. 3 in a cylindrical geometry, is qualitatively similar to the effect conjectured by \\cite{lam07} (hereafter L07). Nevertheless, their investigation about the instability mechanism led them to a different interpretation of the roles of the acoustic and advective-acoustic cycles. We must point out a fundamental difference between the method of L07 and ours: by using a WKB approximation, we have carefuly defined the range of validity of our method, namely short wavelength modes. This guarantees that the advective-acoustic interpretation of the instability mechanism is physical and robust, at least in some parameter range. In contrast, the existence of a purely acoustic instability is still a conjecture because the domain of validity of the method used by L07 is ambiguous: their analytical derivation of a dispersion relation when advection is included requires to neglect terms of order $(v_r /\\omega r)$ while terms of order ${\\cal M}$ are retained. This approximation is not supported by the results of their Fig.~2, which indicates that $(v_{r,{\\rm sh}}/\\omega r_{\\rm sh})$ is comparable to or larger than ${\\cal M}_{\\rm sh}$ for the modes $l=0$ and $l=1$. An accurate description of this acoustic mode, even in a simplified set up, would be useful to gain confidence in its possible existence. In addition to the question of the validity of the approximations used by L07, we find that our results invalidate their reasoning concerning the instability mechanism. They proposed that the advective-acoustic mechanism would be essential if $r_{\\rm sh}/r_*\\ge 10$, whereas a purely acoustic unstable process would be dominant for small shock radii, and they argued that rotation is a key ingredient to discriminate between the two mechanisms. When rotation is included, its effect on SASI has been attributed by L07 to a purely acoustic mechanism, despite the results of their Table 3. However, their view that rotation cannot possibly enhance the growth of the advective-acoustic cycle is clearly incorrect, at least for the short wavelength modes (our Fig.~\\ref{fig4}). \\subsection{Consequences of rotation on supernova explosions} \\label{consequence} The perturbative study of a simple cylindrical configuration has enabled us to cover a large parameter space of shock radii and rotation rates, in order to (i) demonstrate the linear selection of non-axisymmetric modes, (ii) establish a correlation between the preferred direction of the spiral SASI and the rotation of the collapsing core, (iii) identify the advective-acoustic mechanism at work for short wavelength spiral perturbations. The fact that rotation favours a spiral mode $m=1,2$ in a cylindrical flow seems directly connected to the property observed by \\cite{blo07a} in their 3D simulations including rotation. Tracing back the main influence of rotation to the local Doppler shifted frequency $\\omega-m\\Omega$, we may indeed expect a similar destabilization of the spiral modes with a positive value of $m$, a stabilization of the counter-rotating ones, and a comparatively weak influence on the axisymmetric modes. Even a moderate amount of angular momentum results in a shortening of the growth time of SASI through the destabilization of a non-axisymmetric mode. The promising consequences of SASI on both the explosion mechanisms and the pulsar kick could thus be considerably modified, since they were established on the basis of axisymmetric numerical simulations \\citep{bur06,bur07,mar07,sch04,sch06}. Our study suggests that the effect of rotation on the linear phase of SASI can be safely neglected only for slowly rotating progenitors with a specific angular momentum $L\\ll 2\\pi \\cdot 10^{14}$ cm$^2/$s. Although a fast growth of SASI might be helpful to an early shock revival, the dynamical effects of a spiral mode $m=1$, and even $m=2$, on the possible explosion mechanisms are not known yet. If the direction of the kick were determined by the geometry of the most unstable $l=1$ SASI mode, our perturbative approach would suggest a kick-spin misalignment. The strength of the equatorial kick may be diminished by the domination of a symmetric mode $m=2$. It is worth noting however that the relationship between the timescale of the most unstable SASI mode and the onset of explosion is not straightforward, and should be evaluated by future 3D numerical simulations. Our linear approach modestly aims at guiding our intuition for the interpretation of these simulations. \\appendix" }, "0710/0710.3780_arXiv.txt": { "abstract": "Imaging data from the Sloan Digital Sky Survey are used to characterize the population of galaxies in groups and clusters detected with the MaxBCG algorithm. We investigate the dependence of Brightest Cluster Galaxy (BCG) luminosity, and the distributions of satellite galaxy luminosity and satellite color, on cluster properties over the redshift range $0.1 \\le z \\le 0.3$. The size of the dataset allows us to make measurements in many bins of cluster richness, radius and redshift. We find that, within \\rtwo\\ of clusters with mass above $3 \\times 10^{13}$\\Msun, the luminosity function of both red and blue satellites is only weakly dependent on richness. We further find that the shape of the satellite luminosity function does not depend on cluster-centric distance for magnitudes brighter than \\Mi\\ $= -19$. However, the mix of faint red and blue galaxies changes dramatically. The satellite red fraction is dependent on cluster-centric distance, galaxy luminosity and cluster mass, and also increases by $\\sim$5\\% between redshifts 0.28 and 0.2, independent of richness. We find that BCG luminosity is tightly correlated with cluster richness, scaling as $L_{BCG} \\sim M_{200}^{0.3}$, and has a Gaussian distribution at fixed richness, with $\\sigma_{log L} \\sim 0.17$ for massive clusters. The ratios of BCG luminosity to total cluster luminosity and characteristic satellite luminosity scale strongly with cluster richness: in richer systems, BCGs contribute a smaller fraction of the total light, but are brighter compared to typical satellites. This study demonstrates the power of cross-correlation techniques for measuring galaxy populations in purely photometric data. ", "introduction": "\\label{sec:intro} Clusters of galaxies are important systems for studying both galaxy evolution and cosmology. Used as laboratories with well-defined environments, these massive objects are a tool for investigating processes that influence galaxies' physical characteristics. Used as tracers of the underlying mass distribution, they are a tool for investigating the evolution of structure and the nature of dark energy in the Universe. These objectives are closely linked: cosmological studies require accurate knowledge of cluster selection and redshift- and mass-observable relations, and these facts are directly related the evolutionary properties of the cluster galaxy population. In the context of galaxy evolution, the high-density environment of galaxy clusters is a particularly interesting place to examine the galaxy population. Several studies have suggested substantial galaxy transformation in such environments. Galaxy morphology, star-formation rate, and luminosity have long been known to depend on cluster properties and to depart significantly from the cosmological average \\citep[\\eg,][]{Hubble26, Abell62, Oemler74, Dressler80}. Historically, the cluster galaxy content has been quantified by the luminosity function (LF) and type fraction (such as the late-type or blue fraction). Measurement of these quantities as a function of cluster mass, redshift, and distance from the cluster center provides insight into the underlying physical mechanisms responsible for these trends. While the LF is primarily a decreasing function of luminosity, galaxies are bimodal in color and spectral type, with red, early-type galaxies displaying little ongoing star formation, and blue, late-type galaxies exhibiting signs of recent star formation \\citep[\\eg,][]{Strateva01, Baldry04, Bell04, Menanteau06, Blanton05b}. This bimodality was in place by $z\\sim1$ at the latest and may provide a signpost of galaxy transformation \\citep[\\eg,][]{Faber07}. Although this bimodality persists in all environments, the fraction of galaxies in each class (a.k.a. the red and blue fractions) changes systematically with local density; this trend is the so-called morphology-density relationship \\citep{Oemler74,Dressler80,Dressler97,Smith05}. In recent large galaxy surveys this work has been extended to show that a wide range of galaxy properties, including morphology, star-formation rate, and color, depend on local density \\citep{Gomez03,Balogh04a,Balogh04b,Hogg04,Kauffmann04,Tovmassian04,Blanton05b,Christlein05,Croton05,Rojas05,Cooper06,Mandelbaum06,Cooper07}. The advent of large galaxy surveys has made it possible to place observational constraints on both the type fraction and the LF as a function of cluster mass (the conditional luminosity function, CLF), and to further investigate how these quantities depend on other variables. In the Sloan Digital Sky Survey \\citep*[][SDSS]{York00}, \\citet{Goto02} and \\citet{Hansen05} examined the LF as a function of cluster richness and cluster-centric distance for systems found in the SDSS Early Data Release, and \\citet{Weinmann06b} measured the CLF measured from a group catalog derived from the spectroscopic sample of SDSS DR2. Using the 2dF Galaxy Redshift Survey, \\citet{depropris03} and \\citet{Robotham06} compared the LF in high- and low-mass systems; a similar study was performed with a sample of 93 X-ray selected clusters \\citep{LMS04}. The dependence of the LF on galaxy color has been recently investigated \\citep{PopessoLF} as has the galaxy type fraction \\citep{Goto03,depropris04} for clusters in these large surveys. The type fraction of galaxies in clusters depends on both cluster richness and redshift: the fraction of star-forming galaxies at fixed local density is larger at higher redshift, an effect known as the Butcher--Oemler effect \\citep{BO78,ButcherOemler}. This effect is now well-documented, if not entirely well-explained, in clusters over a wide range of masses by studies of the blue fraction, or its converse, the red fraction \\citep{Rakos95,Margoniner00,Ellingson01,KodamaBower01,Margoniner01,depropris04,Martinez06,Gerke07}. Other indicators of galaxy state, including galaxy morphology and emission line strength, also show a trend with redshift \\citep{AS93,ODB97,Balogh97,Couch98,vanDokkum00,Fasano00,Lubin02,Goto03,Treu03,Wilman05,Poggianti06,Desai07,vanderwel07}. However, few samples to date have had both large numbers of systems and well-understood mass proxies, so the dynamical range and mass resolution of these previous works has been somewhat limited. Another characteristic of galaxy clusters is the presence of a highly-luminous galaxy near the cluster center --- the Brightest Cluster Galaxy (BCG). In addition to being extraordinarily luminous, BCGs differ in a number of ways from other cluster members: they tend to have extended light profiles \\citep{Matthews64, Tonry87, Schombert88, Gonzalez00, Gonzalez03}, larger size at fixed luminosity than other early types \\citep[][and references therein]{Bernardi07} and may contain a larger fraction of dark matter than typical galaxies \\citep[\\eg,][]{Mandelbaum06,Anja07}. Also, while the traditional fitting function to the LF of \\citet{Schechter76} provides a good fit for satellite galaxies, BCGs follow a different distribution, causing the so-called ``bright end bump'' \\citep[\\eg,][]{Hansen05}. This difference between the BCG and other satellites in a cluster is also manifested in the luminosity gap statistic, the difference in luminosity between the BCG and the next brightest cluster member, which may be indicative of the special accretion history of BCGs \\citep{Ostriker75, Tremaine77, Loh06}. BCGs are also distinct from other galaxies with similar mass that are not at the center of cluster-sized potential wells \\citep{Anja07}. Indeed, the properties of BCGs seem to be closely linked to properties of their host clusters \\citep{Sandage73,Sch83,Schombert88, Edge91, Brough02, Degrandi04,Brough05,Loh06}, including to the masses of their parent halos \\citep{LinMohr04, Mandelbaum06, ZCZ07}. The outer light profile of BCGs often merges smoothly with the diffuse intra-cluster light (ICL), suggesting again a coupling between the BCG and the host halo \\citep{Gonzalez05}. Generally the BCG+ICL light is closely linked with cluster mass \\citep{Zibetti05,Conroy07ICL,Purcell07,Gonzalez07}. As BCGs have properties different from those of the rest of the cluster members, BCGs and satellites are often analyzed separately, and we follow this convention here. It is expected that the properties of cluster galaxies are closely tied to the merging and accretion history of their parent dark matter halos. Current models based on Cold Dark Matter (CDM) suggest that the stars in BCGs were formed in dense peaks quite early but that the BCGs were assembled in a series of galaxy merging events that continue until relatively recent times \\citep[e.g.][]{AS98,Dubinski98,Gao04a,BK06,DeLucia07}. Satellite galaxies in clusters are now generally understood to be hosted by smaller dark matter halos that have merged into the parent halo. Several features of the observed cluster galaxy populations can be understood based on the assembly history of the parent halo \\citep[\\eg,][]{Poggianti06, Iro07}. One such characterization is the core distinction between central galaxies and satellites: the central concentrations of mass and light in massive halos continue to build up while the growth of satellite systems is halted upon accretion. BCG luminosity, and the correlations between this luminosity and both cluster mass and satellite properties, can thus provide insight into the assembly histories of clusters. The merger history of dark matter halos alone is not enough to account for the bimodality in galaxy properties and its dependence on redshift and environment. Several physical processes have been proposed to transform star-forming galaxies into the typical cluster galaxies on the red sequence. Although the relative strengths of these processes are still hotly debated, a consensus is emerging that the main transformation mechanisms are related to the mass of the host halo and whether (and for how long) the galaxy has been a satellite within a larger system. Among the suggested transformation mechanisms, some are expected to be most effective in rich clusters, such as ram-pressure stripping \\citep{GunnGott72}, interaction with the cluster potential \\citep{ByrdValtonen90} and high-velocity close encounters \\citep[``harassment;''][]{Moore96}. However, studies of very poor systems have shown that the environmental dependence of galaxy properties is not limited to the richest objects \\citep[\\eg,][]{Zabludoff98,Weinmann06a,Gerke07}. Processes that can operate efficiently in low velocity dispersion systems therefore must also play a role in shaping the galaxy population: \\eg, galaxy mergers \\citep{TT72} and ``strangulation,'' a cutoff to gas accretion onto galaxy disks by stripping or AGN feedback \\citep{Larson80,BNM00,Croton06}. It is likely that some combination of these effects is at work. Distinguishing their relative significance requires precisely quantifying cluster galaxy properties over a wide range of masses and as a function of cluster-centric distance. These data will provide information on both the assembly histories of clusters and on the physical mechanisms that trigger and quench star formation. Models of galaxy evolution in a cosmological context most readily predict galaxy properties as a function of halo mass rather than cluster observables. In order to make these comparisons, a reliable mass--observable relationship is a prerequisite. In addition, there is consensus that the luminosity function and type fraction both depend on a number of variables, complicating detailed comparison between clusters of different mass. For example, the cluster galaxy LF depends on cluster-centric distance, and the size of the bound regions of clusters scales with mass. In order to make physically meaningful comparisons between LFs of different mass clusters, an aperture scaled to the bound region is therefore preferable to a fixed metric aperture. With recent extensive surveys providing well-calibrated cluster catalogs spanning a wide range in mass, it is possible to examine in detail the dependence of the cluster galaxy population on several cluster and galaxy properties simultaneously. Large, homogeneous photometric surveys such as the SDSS provide rich data with which to characterize the cluster galaxy population. These data have been used to define large, robust, clean samples of galaxy clusters with accurate photometric redshifts. These samples are sizable enough to split on several variables allowing detailed statistical exploration of the galaxy populations in clusters. Currently, the largest sample of clusters available is the \\maxbcg\\ catalog from the Sloan Digital Sky Survey \\citep{Koester07a}. The selection effects of the cluster-finding algorithm are well understood \\citep{Koester07b, Rozo07a}, and there are a number of studies exploring the mass--richness relationship for these objects \\citep{Rozo07b,Becker07,Sheldon07a, Johnston07b,Rykoff07}. In this work, we convert cluster richness to cluster mass using weak lensing measurements of the \\maxbcg\\ mass--richness relation \\citep{Sheldon07a,Johnston07b}. The quality and quantity of these data allow for detailed measurements of a variety of cluster galaxy properties as a function of cluster mass and radius. For satellite galaxies we then measure the luminosity function of all, red, and blue satellites conditional on both mass and cluster radius, and investigate the dependence of the red fraction of satellites on cluster mass, redshift, galaxy luminosity and distance from cluster center; for BCGs we quantify the dependence on cluster mass of both the BCG luminosity and the relationship between the BCG luminosity and satellite galaxy luminosities. Although this sample of clusters extends only to $z = 0.3$, these objects provide a valuable low-redshift baseline with which higher redshift samples may be compared. In this paper we use a statistical background-subtraction technique to measure mean galaxy properties over a wide range of color and luminosity in \\maxbcg\\ clusters. We average the signal from many clusters binned by cluster properties, and statistically subtract the contribution from random galaxies along the line of sight. This method of cross-correlating clusters with the galaxy population provides very precise statistical measurements, and allows us to study blue and low luminosity galaxies that are indistinguishable from the background in individual clusters. We test the background-correction algorithm by running the full analysis on realistic mock catalogs, and find that we are able to robustly recover 3D cluster properties using these methods. The statistical techniques presented here for background correction of photometric data will be directly applicable to future large multi-band imaging programs, including the Dark Energy Survey \\citep[DES\\footnote{\\tt http://www.darkenergysurvey.org},][]{DES} and the Large Synoptic Survey Telescope \\citep[LSST\\footnote{{\\tt http://www.lsst.org}},][]{LSST} and the Panoramic Survey Telescope \\& Rapid Response System \\citep[Pan-STARRS\\footnote{{\\tt http://pan-starrs.ifa.hawaii.edu}},][]{PANSTARRS} and will be essential for leveraging these data to provide the desired insights into cosmology and galaxy evolution. The paper is organized as follows: in \\S\\ \\ref{sec:data} we describe the SDSS and simulation data used; we present the stacking and background-correction method in \\S\\ \\ref{sec:methods}. Our primary results are given in \\S\\ \\ref{sec:results}: \\S\\ \\ref{sec:sats} presents the luminosity and color characteristics of the satellite population, while \\S\\ \\ref{sec:BCGs} discusses the BCG population. A summary and discussion of the implications of the results is given in \\S\\ \\ref{sec:conclusion}. The notation used for cluster-related variables in previous \\maxbcg\\ work includes defining \\Ntwo\\ and $L_{200}$ as the counts and $i$-band luminosity of red-sequence galaxies within the measurement aperture of the cluster finder and with L $> 0.4L_*$. We note that as the aperture for cluster finding was determined with a previous definition of richness, it is not strictly the true value of \\rtwo\\ for these systems (in fact, it is larger), and only red galaxies are included in these definitions. In this work we will refer to the total excess luminosity associated with the light from galaxies of {\\em all} colors above a luminosity threshold and within the measured \\rtwo\\ of these systems as $L_{200}$. In addition, we follow the standard convention of using $R$ to to denote projected, 2D radii and $r$ to refer to deprojected, 3D radii. Where necessary for computing distances, we assume a flat, LCDM cosmology with H$_{0} = 100h$ km s$^{-1}$ Mpc$^{-1}$, $h = 0.7$, and matter density $\\Omega_m = 0.3$. ", "conclusions": "\\label{sec:conclusion} In this study, we have examined the properties of cluster-associated galaxies, separating BCGs from satellites and focusing on trends in galaxy color and luminosity as a function of cluster richness, distance from cluster center, and redshift. We use clusters in the SDSS \\maxbcg\\ sample, the largest set of clusters identified to date. Employing photometric data alone, we apply cross-correlation background-correction techniques to characterize the cluster-associated galaxy population, within $5\\times$\\rtwo\\ and brighter than \\Mi\\ $< -19$, around \\nclust\\ systems spanning more than two decades in mass in the redshift range $0.1 \\le z \\le 0.3$. Our principle results are as follows. \\begin{enumerate} \\item The luminosity function of satellites within \\rtwo\\ as a function of cluster mass for systems with mass greater than $3 \\times 10^{13}$\\Msun\\ shows remarkable uniformity for \\Mi\\ $< -19$. The characteristic satellite luminosity \\Lstar\\ is only weakly dependent on cluster richness. \\item The shape of the luminosity function of satellites brighter than \\Mi\\ $< -19$ does not change with cluster-centric radius. However, the color-separated luminosity functions of satellites as a function of \\rad\\ and of \\Ntwo\\ show that the mix of sub-\\Lstar\\ red and blue galaxies changes dramatically as a function of radius. In contrast, the relative number of red and blue galaxies at the bright end is roughly constant with radius. \\item The average color of satellite galaxies is redder near cluster centers, but this trend is largely a reflection of the changing ratio of red to blue galaxies mentioned above. This effect is a very weak function of cluster mass over the range investigated here. \\item The fraction of red galaxies increases with cluster mass over the full range explored, although only weakly for \\Mtwo $\\gae 10^{14}$\\Msun. This fraction decreases with cluster-centric distance until $2\\times$\\rtwo, decreases with luminosity for $L<$ \\Lstar, and is constant for $r>2\\times$\\rtwo\\ and $L>$ \\Lstar. \\item The fraction of cluster galaxies within \\rtwo\\ that are red increases by $\\sim$5\\% during the 0.8 Gyr between redshift $z=0.28$ to $z=0.2$; this change is roughly independent of mass over the range investigated. \\item The luminosity of BCGs and the ratios of BCG luminosity to total cluster luminosity and to characteristic satellite luminosity are all correlated strongly with cluster mass, and we have quantified each of these scalings over the mass range $3 \\times 10^{13}$\\Msun\\ to $9 \\times 10^{14}$\\Msun. The BCG luminosity has a Gaussian distribution at fixed cluster richness, with dispersion $\\sigma_{logL} \\sim 0.17$ for clusters with \\Mtwo\\ $ > 10^{14}$\\Msun. \\end{enumerate} While these results are in general agreement with previous observational work, due to the volume probed we are able to investigate a wider mass range, extending the statistics to higher mass than previous samples. We are also able to split the sample into finer bins for several variables than has previously been possible. The \\maxbcg\\ selection function, which is well-understood for most of the richness range used here, is less well quantified for clusters with \\Ntwo\\ $< 10$. This low richness set of clusters may be less complete and pure, which could result in the significant difference in LF shape as compared to higher \\Ntwo\\ systems, or contribute to the drop-off in \\fred\\ for low richness systems. However, not all cluster galaxy properties change significantly at \\Ntwo\\ $= 10$ (\\eg, \\Lbcg/\\Ltwo) and there is no break at a particular \\Ntwo\\ in either the mass--observable relationship \\citep{Johnston07b} or mass-to-light ratios \\citep{Sheldon07b} as measured by lensing. Interestingly, it is the quantities that most closely trace the total mass of the systems (\\ie, \\Lbcg, \\Ltwo and the lensing signal) that are smoothly scaling over the full richness range, while quantities that are related to the mix of galaxies within clusters (\\ie, the LF and \\fred) are the ones that change more dramatically. We hypothesize that, at these low richnesses, \\maxbcg\\ is finding legitimate low-mass systems, but that the selection priors demanding the close proximity of only a few red galaxies result in finding only systems with the observed mix of galaxies and not necessarily {\\em all} systems of this low mass. Further investigation using lower mass threshold mock catalogs is needed to understand in detail the selection of low-richness systems. Our results can shed light both on the processes that build up the galaxy population in clusters and distinguish central galaxies from satellites, as well as the processes that are responsible for the galaxy transformation from blue to red. With respect to the former, the results presented here fit well within the basic picture of galaxy formation in CDM: that galaxy properties are likely linked to the formation history of a cluster's dark matter halo and its substructures. Although detailed comparisons are beyond the scope of this work, our results qualitatively match both HOD constraints from clustering statistics as well as models based on matching the abundance of halos and subhalos to galaxies. The HOD framework provides a way to examine the galaxy population in both the observed Universe and in models of galaxy formation and evolution. Without needing to identify specific groups or clusters in the data, halo model interpretations of the statistics of luminosity-dependent galaxy clustering result in specific predictions for trends of both central and satellite galaxies as a function of halo mass that are in reasonable agreement with the findings presented here, especially for the relationship between central and satellite luminosities \\citep[\\eg,][]{Berlind03,Skibba07}. Models which link the properties of galaxies directly to their halos and subhalos also agree broadly with several of the results presented here \\citep[\\eg,][]{ Conroy06, VO07}. Detailed predictions for several cluster statistics from such a model will be presented in \\citet{Iro07}. Our results on scatter in the BCG luminosity at fixed cluster richness very likely provide an upper limit on scatter in central galaxy luminosity at fixed mass, as the scatter between halo mass and cluster richness should act to increase this scatter. The fact that this scatter is already fairly small provides further support for the tight coupling between halo mass and galaxy luminosity that is the basis of these models. In addition to processes that cause physical changes to satellites resulting from interactions with the cluster gas, cluster potential, or other satellites, presumably some of the satellites are lost due to being accreted onto the BCG. In detail, this process likely results in the stellar component of the disrupted galaxy joining both the BCG and ICL \\citep{Conroy07ICL}, but nonetheless should result in a BCG population closely linked to both halo mass and satellite population, as is observed. Indeed, \\Lbcg/\\Ltwo\\ and \\Lbcg/\\Lstarsat\\ must be intimately related to the processes responsible for BCG growth. That BCGs get brighter as a function of cluster mass faster than do typical satellies may be further evidence that BCGs are different (in their merger history) than typical satellites. Understanding the timescales and mechanisms for galaxies to transform from star-forming galaxies onto the red sequence is one of the primary current challenges for galaxy formation theories. There are several processes that can operate within clusters to shape the population of the cluster galaxies, such as ram pressure stripping, harassment and strangulation, that directly influence the galaxies' gas content and thus their subsequent star formation \\citep[for a recent review, see][]{DeLucia06}. Clearly many of the processes responsible for this transformation are related either to the mass of the host halo or how long the galaxies have been satellites. The relative strengths of various effects are still rather uncertain, however. What fraction of galaxies become red while they are central galaxies? Do the processes happen only for satellites in a certain mass range, or do they happen equally for all satellite galaxies? On what timescales do these processes operate? Our measurements of how the red fraction scales with cluster mass, radius and redshift will be instrumental in answering these questions. We address a few of these issues here. Ram pressure stripping predicts that \\fred\\ will be inversely correlated with cluster-centric distance, and larger for both brighter galaxies and more massive halos. However, our results indicate that \\fred\\ is essentially independent of galaxy luminosity at fixed \\Ntwo\\ for galaxies brighter than \\Lstar\\ and that the radial trend in \\fred\\ is not any more pronounced in high \\Ntwo\\ systems. These results indicate that ram pressure stripping cannot be the dominant process at work to transform the galaxy population. The harassment scenario predicts that \\fred\\ will be anticorrelated with cluster mass, as is observed; however, if this mechanism were dominant, then at fixed cluster mass \\fred\\ would be expected to be larger for less luminous galaxies \\citep{Weinmann06a}, contradicting the observed trends. Strangulation, where star formation in infalling galaxies is halted because no further gas accretion is allowed, makes several predictions that are in good agreement with our observations. For example, the model presented by \\citet{Diaferio01} predicts: that the mean satellite color gets bluer as a function of cluster-centric distance until reaching a plateau at the field value around 2--3\\rtwo; that the mean satellite color depends on halo mass only for $M \\lae 5 \\times 10^{13}$\\Msun; and that the incidence of blue galaxies in clusters increases at higher redshift. However, this model also predicts that both bulge- and disk-dominated satellies will get redder toward the central regions of clusters, a trend that we do not see in our red/blue sample split (or course, our color-separated subsamples are not directly comparable to their morphology-separated samples). Note that the predictions of such a model will in detail depend on its implementation. Our observation that the satellite red fraction changes in the same way with redshift regardless of cluster richness is significant evidence that the timescale of the physics responsible for quenching is the same in systems over the full mass range examined. The difference in cosmic time between the median redshifts of our two samples is approximately the dynamical timescale, over which we observe a $\\sim$5\\% change in \\fred, and this observation should allow for useful constraints in studies on the details of quenching mechanisms. A consensus is emerging from a variety of modeling efforts that star formation proceeds most efficiently in a mass range around $L_*$, and is less efficient in more massive halos and for satellite galaxies. Simple models based on this type of assumption can produce many of the rough trends that we have seen here; for example, that the mean color of galaxies will not change significantly as a function of halo mass for the range of masses we investigate here \\citep[\\eg,][]{Diaferio01}, and that the galaxy type fraction should be a weak function of halo mass for $M$ \\gae $10^{13}$\\Msun\\ \\citep{Berlind03,Zheng05,Cooray05}. Furthermore, the general concept of galaxies falling in, quenching, and fading matches well with the observed radial trends of LF and red fraction. However, most of the detailed semi-analytic modeling efforts have had some trouble matching in detail observations of the color distribution of galaxies and how it changes with environment \\citep[see \\eg,][]{Coil07}. Unresolved issues include the rates at which star formation is triggered or shut off, and accordingly red galaxies tend to be overproduced. To understand in detail the physical mechanisms responsible for the quenching of star formation as clusters are assembled, further work is needed to accurately reproduce the observed trends in the cluster galaxy population. The present findings set a local-universe target for modeling results, and provide some guidance for the relative importance of some of the germane effects. A substantial effort over the next decade will be devoted to large-scale, multi-band photometric surveys, including DES, Pan-STARRS, and LSST. Although the primary science driver of many of these projects is to investigate the nature of dark energy, the resulting data are likely to also provide strong constraints on the processes of galaxy evolution. The results presented here provide a low-redshift baseline against which current and future high-redshift samples may be compared. From a technical standpoint, these data are informative to next-generation cluster-finding techniques, and are useful input for creating the mock catalogs necessary for interpreting cluster surveys. Furthermore, the techniques presented here, which use photometric data alone, are directly applicable to these upcoming imaging surveys, and will thus enable detailed studies of the galaxy population at significantly higher redshifts without extensive spectroscopy." }, "0710/0710.1099_arXiv.txt": { "abstract": "The reliability of quiet Sun magnetic field diagnostics based on the \\ion{Fe}{1} lines at 6302 \\AA \\, has been questioned by recent work. We present here the results of a thorough study of high-resolution multi-line observations taken with the new spectro-polarimeter SPINOR, comprising the 5250 and 6302 \\AA \\, spectral domains. The observations were analyzed using several inversion algorithms, including Milne-Eddington, LTE with 1 and 2 components, and MISMA codes. We find that the line-ratio technique applied to the 5250~\\AA \\, lines is not sufficiently reliable to provide a direct magnetic diagnostic in the presence of thermal fluctuations and variable line broadening. In general, one needs to resort to inversion algorithms, ideally with realistic magneto-hydrodynamical constrains. When this is done, the 5250~\\AA \\, lines do not seem to provide any significant advantage over those at 6302~\\AA . In fact, our results point towards a better performance with the latter (in the presence of turbulent line broadening). In any case, for very weak flux concentrations, neither spectral region alone provides sufficient constraints to fully disentangle the intrinsic field strengths. Instead, we advocate for a combined analysis of both spectral ranges, which yields a better determination of the quiet Sun magnetic properties. Finally, we propose the use of two other \\ion{Fe}{1} lines (at 4122 and 9000~\\AA ) with identical line opacities that seem to work much better than the others. ", "introduction": "\\label{sec:intro} The empirical investigation of quiet Sun\\footnote{In this work, we use the term ``quiet Sun'' to refer to the solar surface away from sunspots and active regions.} magnetism is a very important but extremely challenging problem. A large (probably dominant) fraction of the solar magnetic flux resides in magnetic accumulations outside active regions, forming network and inter-network patches (e.g., \\citeNP{SNSA02}). It is difficult to obtain conclusive observations of these structures, mainly because of two reasons. First, the size of the magnetic concentrations is much smaller than the spatial resolution capability of modern spectro-polarimetric instrumentation. Estimates obtained with inversion codes yield typical values for the filling factor of the resolution element between $\\sim$1\\% and 30\\% . The interpretation of the polarization signal becomes non-trivial in these conditions and one needs to make use of detailed inversion codes to infer the magnetic field in the atmosphere. Second, the observed signals are extremely weak (typically below $\\sim$1\\% of the average continuum intensity), demanding both high sensitivity and high resolution. Linear polarization is rarely observed in visible lines, so one is usually left with Stokes~$I$ and~$V$ alone. \\citeN{S73} proposed to use the pair of \\ion{Fe}{1} lines at 5247 and~5250~\\AA \\, which have very similar excitation potentials and oscillator strengths (and, therefore, very similar opacities) but different Land\\'e factors, to determine the intrinsic field strength directly from the Stokes~$V$ line ratio. That work led to the subsequent popularization of this spectral region for further studies of unresolved solar magnetic structures. Later, the pair of \\ion{Fe}{1} lines at 6302~\\AA \\, became the primary target of the Advanced Stokes Polarimeter (ASP, \\citeNP{ELT+92}), mainly due to their lower sensitivity to temperature fluctuations. The success of the ASP has contributed largely to the currently widespread use of the 6302~\\AA \\, lines by the solar community. Recent advances in infrared spectro-polarimetric instrumentation now permit the routine observation of another very interesting pair of \\ion{Fe}{1} lines, namely those at 15648 and 15653~\\AA \\, (hereafter, the 1.56~$\\mu$m lines). Examples are the works of \\citeN{LR99}; \\citeN{KCS+03}. The large Land\\' e factors of these lines, combined with their very long wavelengths, result in an extraordinary Zeeman sensitivity. Their Stokes~$V$ profiles exhibit patterns where the $\\sigma$-components are completely split for fields stronger than $\\sim$400~G at typical photospheric conditions. They also produce stronger linear polarization. On the downside, this spectral range is accesible to very few polarimeters. Furthermore, the 1.56~$\\mu$m lines are rather weak in comparison with the above-mentioned visible lines. Unfortunately, the picture revealed by the new infrared data often differs drastically from what was being inferred from the 6302~\\AA \\, observations (e.g., \\citeNP{LR99}; \\citeNP{KCS+03}; \\citeNP{SNSA02}; \\citeNP{SNL04}; \\citeNP{DCSAK03}), particularly in the inter-network. \\citeN{SNSA03} proposed that the discrepancy in the field strengths inferred from the visible and infrared lines may be explained by magnetic inhomogeneities within the resolution element (typically 1\\arcsec). If multiple field strengths coexist in the observed pixel, then the infrared lines will be more sensitive to the weaker fields of the distribution whereas the visible lines will provide information on the stronger fields (see also the discussion about polarimetric signal increase in the 1.56 $\\mu$m lines with weakening fields in \\citeNP{SAL00}). This conjecture has been tested recently by \\citeN{DCSAK06} who modeled simultaenous observations of visible and infrared lines using unresolved magnetic inhomogeneities. A recent paper describing numerical simulations by \\citeN{MGCRC06} casts some doubts on the results obtained using the 6302~\\AA \\, lines. Our motivation for the present work is to resolve this issue by observing simultaneously the quiet Sun at 5250 and 6302~\\AA . We know that unresolved magnetic structure might result in different field determinations in the visible and the infrared, but the lines analyzed in this work are close enough in wavelengths and Zeeman sensitivities that one would expect to obtain the same results for both spectral regions. ", "conclusions": "\\label{sec:conc} The ratio of Stokes~$V$ amplitudes at 5250 and 5247~\\AA \\, is a very good indicator of the intrinsic field strength in the absence of line broadening, e.g. due to turbulence. However, line broadening tends to smear out spectral features and reduce the Stokes~$V$ amplitudes. This reduction is not the same for both lines, depending on the profile shape. If the broadening could somehow be held constant, one would obtain a line-ratio calibration with very low scatter. However, if the broadening is allowed to fluctuate, even with amplitudes as small as 1~km~s$^{-1}$, the scatter becomes very large. Fluctuations in the thermal conditions of the atmosphere further complicate the analysis. This paper is not intended to question the historical merits of the line-ratio technique, which led researchers to learn that fields seen in the quiet Sun at low spatial resolution are mostly of kG strength with small filling factors. However, it is important to know its limitations. Otherwise, the interpretation of data such as those in Figure~\\ref{fig:mapratios} could be misleading. Before this work, most of the authors were under the impression that measuring the line ratio of the 5250~\\AA \\, lines would always provide an accurate determination of the intrinsic field strength. With very high-resolution observations, such as those expected from the Advanced Technology Solar Telescope (ATST, \\citeNP{KRK+03}) or the Hinode satellite, there is some hope that most of the turbulent velocity fields may be resolved. In that case, the turbulent broadening would be negligible and the line-ratio technique would be more robust. However, even with the highest possible spatial resolution, velocity and temperature fluctuations along the line of sight will still produce turbulent broadening. From the study presented here we conclude that, away from active region flux concentrations, it is not straightforward to measure intrinsic field strengths from either 5250 or 6302~\\AA \\, observations taken separately. Weak-flux internetwork observations would be even more challenging, as demonstrated recently by \\citeN{MG07}. Surprisingly enough, the 6302~\\AA \\, pair of \\ion{Fe}{1} lines is more robust than the 5250~\\AA \\, lines in the sense that it is indeed possible to discriminate between weak and strong field solutions if one is able to rule out a thermal stratification with temperatures that increase outwards. Even so, this is only possible when one employs an inversion code that has sufficient MHD constrains (an example is the MISMA implementation used here) to reduce the space of possible solutions. The longitudinal flux density obtained from inversions of the 6302~\\AA \\, lines is better determined than those obtained with 5250~\\AA . This happens regardless of the inversion method employed, although using a code like LILIA provides better results than a simpler one such as MELANIE. The best fits to average network profiles correspond to strong kG fields, as one would expect. An interesting conclusion of this study is that it is possible to obtain reliable results by inverting simultaneous observations at both 5250 and 6302~\\AA . Obviously this would be possible with relatively sophisticated algorithms (e.g., LTE inversions) but not with simple Milne-Eddington inversions. The combination of two other \\ion{Fe}{1} lines, namely those at 4122 and 9000~\\AA , seems to provide a much more robust determination of the quiet Sun magnetic fields. Unfortunately, these lines are very distant in wavelength and few spectro-polarimeters are capable of observing them simultaneously. Examples of instrument with this capability are the currently operational SPINOR and THEMIS, as well as the planned ATST and GREGOR. Depending on the evolution time scales of the structures analyzed it may be possible for some other instruments to observe the blue and red lines alternatively." }, "0710/0710.3325_arXiv.txt": { "abstract": "{} {In this paper we study the possibility of testing $CPT$ symmetry with Cosmic Microwave Background (CMB) measurements.} {Working with an effective lagrangian of the photon with $CPT$ violation ${\\cal L} \\sim p_{\\mu}A_{\\nu}\\tilde F^{\\mu\\nu}$ which causes the polarization vectors of the propagating CMB photons rotated, we determine the rotation angle $\\Delta\\alpha$ using the BOOMERanG 2003 and the WMAP3 angular power spectra.} {In this analysis we have included the newly released $TC$ and $GC$ ($l<450$) information of WMAP3 and found $\\Delta\\alpha=-6.2\\pm3.8$ deg at $68\\%$ confidence level.} {This result increases slightly the significance for the $CPT$ violation obtained in our previous paper (Feng et al. 2006) $\\Delta\\alpha=-6.0 \\pm 4.0$ deg (1$\\sigma$). Furthermore we examine the constraint on the rotation angle from the simulated polarization data with Planck precision. Our results show that the future Planck measurement will be sensitive to $\\Delta \\alpha$ at the level of $0.057$ deg and able to test the $CPT$ symmetry with a higher precision.} ", "introduction": "\\label{Introduction} In the standard model of particle physics $CPT$ is a fundamental symmetry. Probing its violation is an important way to search for the new physics beyond the standard model. Up to now, $CPT$ symmetry has passed a number of high precision experimental tests and no definite signal of its violation has been observed in the laboratory. So, the present $CPT$ violating effects, if exist, should be very small to be amenable to the experimental limits. The $CPT$ symmetry could be dynamically violated in the expanding universe \\cite{Li:2007}. The cosmological $CPT$ violation mechanism investigated in the literature \\cite{Li:2001st,Li:2002wd,Li:2004hh,Feng:2004mq,Li:2007} has an interesting feature that the $CPT$ violating effects at present time are too small to be detected by the laboratory experiments but large enough in the early universe to account for the generation of matter-antimatter asymmetry \\cite{Li:2001st,Li:2002wd,Li:2004hh,Li:2007}. And more importantly, this type of $CPT$ violating effects could be accumulated to be observable in the cosmological experiments \\cite{Feng:2004mq,Li:2007,Feng:2006dp}. With the accumulation of high quality observational data, especially those from the CMB experiments, cosmological observation becomes a powerful way to test the $CPT$ symmetry. Here we study the CMB polarizations and $CPT$ violation in the photon sector with an effective lagrangian \\cite{Carroll:1989vb,Carroll:1990zs}: \\begin{equation}\\label{Lagrangian} \\mathcal{L} = -\\frac{1}{4}F_{\\mu\\nu}F^{\\mu\\nu}+\\mathcal{L}_{cs}~, \\end{equation} where $\\mathcal{L}_{cs}\\sim p_{\\mu}A_{\\nu}\\tilde F^{\\mu\\nu}$ is a Chern-Simons term, $p_{\\mu}$ is an external vector and $\\tilde F^{\\mu\\nu}=(1/2)\\epsilon^{\\mu\\nu\\rho\\sigma}F_{\\rho\\sigma}$ is the dual of the electromagnetic tensor. This Lagrangian is not gauge invariant, but the action is gauge independent if $\\partial_{\\nu}p_{\\mu}=\\partial_{\\mu}p_{\\nu}$. This may be possible if $p_{\\mu}$ is constant in spacetime or the gradient of a scalar field in the quintessential baryo-/leptogenesis \\cite{Li:2002wd,Li:2001st,quin_baryogenesis} or the gradient of a function of the Ricci scalar in gravitational baryo-/leptogenesis \\cite{Li:2004hh,R}. The Chern-Simons term violates Lorentz and $CPT$ symmetries when the background value of $p_{\\mu}$ is nonzero. One of the physical consequences of the Chern-Simons term is the rotation of the polarization direction of electromagnetic waves propagating over large distances \\cite{Carroll:1989vb}. From the Lagrangian (\\ref{Lagrangian}), we can directly obtain the equation of motion for the electromagnetic field: \\begin{equation}\\label{maxwell1} \\nabla_{\\mu}(\\nabla^{\\mu}A^{\\nu}-\\nabla^{\\nu}A^{\\mu})=-p_{\\mu} \\epsilon^{\\mu\\nu\\rho\\sigma}(\\nabla_{\\rho}A_{\\sigma}-\\nabla_{\\sigma}A_{\\rho})~. \\end{equation} After imposing Lorentz gauge condition $\\nabla_{\\mu}A^{\\mu}=0$, it becomes: \\begin{equation}\\label{maxwell2} \\nabla_{\\mu}\\nabla^{\\mu}A^{\\nu}+R^{\\nu}_{\\mu}A^{\\mu}=-p_{\\mu} \\epsilon^{\\mu\\nu\\rho\\sigma}(\\nabla_{\\rho}A_{\\sigma}-\\nabla_{\\sigma}A_{\\rho})~, \\end{equation} where $R^{\\nu}_{\\mu}$ is the Ricci tensor. With the geometric optics approximation, the solution to the equation of motion is expected to be: $A^{\\mu}={\\rm Re}[(a^{\\mu}+\\epsilon b^{\\mu}+\\epsilon^2 c^{\\mu}+...)e^{iS/\\epsilon}]$, where $\\epsilon$ is a small number. With this ansatz, one can easily see that the Lorentz gauge condition implies $k_{\\mu}a^{\\mu}=0$, where the wave vector $k_{\\mu}\\equiv \\nabla_{\\mu}S$ is orthogonal to the surfaces of constant phase and represents the direction which photons travel along with. The vector $a^{\\mu}$ is the product of a scalar amplitude $A$ and a normalized polarization vector $\\varepsilon^{\\mu}$, $a^{\\mu}=A\\varepsilon^{\\mu}$, with $\\varepsilon_{\\mu}\\varepsilon^{\\mu}=1$. Hence in the Lorentz gauge, the wave vector $k_{\\mu}$ is orthogonal to the polarization vector $\\varepsilon^{\\mu}$. Substituting this solution into the modified Maxwell equation (\\ref{maxwell2}) and neglecting the Ricci tensor we have at the leading order of $\\epsilon$ the equation is $k_{\\mu}k^{\\mu}=0$. It indicates that photons still propagate along the null geodesics. The effect of Chern-Simons term appears at the next order, $k^{\\mu}\\nabla_{\\mu}\\varepsilon^{\\nu}=-p_{\\mu}\\epsilon^{\\mu\\nu\\rho\\sigma} k_{\\rho}\\varepsilon_{\\sigma}$. We can see that the Chern-Simons term makes $k^{\\mu}\\nabla_{\\mu}\\varepsilon^{\\nu}$ not vanished. This means that the polarization vector $\\varepsilon^{\\nu}$ is not parallel transported along the light-ray. It rotates as the photon propagates in spacetime. We consider here the spacetime described by spatially flat Friedmann-Robertson-Walker (FRW) metric. The null geodesics equation is $(k^0)^2-k^ik^i=0$. We assume that photons propagate along the positive direction of $x$ axis, \\emph{i.e.} $k^{\\mu}=(k^0,k^1,0,0)$ and $k^1=k^0$. Gauge invariance guarantees that the polarization vector of the photon has only two independent components which are orthogonal to the propagating direction. So, we are only interested in the changes of the components of the polarization vector, $\\varepsilon^2$ and $\\varepsilon^3$. Assuming $p_{\\mu}=p_0$ to be a non-vanishing constant, we obtain the following equations: $d\\varepsilon^2/d\\lambda+\\mathcal{H}k^0\\varepsilon^2=p_0 k^0\\varepsilon^3$, $d\\varepsilon^3/d\\lambda+\\mathcal{H}k^0\\varepsilon^3= - p_0 k^0\\varepsilon^2$, where we have defined the affine parameter $\\lambda$ which measures the distance along the light-ray, $k^{\\mu}\\equiv dx^{\\mu}/d\\lambda$, and the reduced expansion rate $\\mathcal{H}\\equiv \\dot a/a$. The polarization angle is defined as $\\chi\\equiv\\arctan{(\\varepsilon^3}/{\\varepsilon^2})$. It is easy to find that the rotation angle is \\begin{equation}\\label{kmu1} \\Delta\\chi\\equiv \\chi_0-\\chi_z=-\\int^{\\eta_0}_{\\eta_z} ~p_0 ~d\\eta=\\int^{t_z}_{t_0} p_0 ~\\frac{dt}{a}~, \\end{equation} where the subscript $z$ is the redshift of the source when the light was emitted. For CMB photons, the source is the last scattering surface with $z\\simeq 1100$ \\footnote{Besides the CMB photons which come from the last scattering surface, we might observed the different CMB photons which travelled different distances as well. However, we are only interested in linear perturbations of CMB photons in this paper. CMB polarizations are already linear phenomena. They are not existent at the zeroth order. When calculating the variations of polarizations due to $CPT$ violation in the perturbation theory up to linear order, we may ignore the fluctuations of the travelling distances of CMB photons. Otherwise, these fluctuations combined with polarizations would give higher order result which are beyond the scope of this paper. So, in this paper, we only consider linear perturbations and we can assume that each CMB photon detected by us travelled the same distance.}. The subscript $0$ indicates the present time. As we know, a vector rotated by an angle $\\Delta\\chi$ in a fixed coordinates frame is equivalent to a fixed vector observed in a coordinates frame which is rotated by $-\\Delta\\chi$. So, with the notion of coordinates frame rotation, the rotation angle is \\begin{equation}\\label{kmu2} \\Delta\\alpha=-\\Delta\\chi=\\int^{t_0}_{t_z} ~ p_0 ~ \\frac{dt}{a}~ = p_0 r_z~. \\end{equation} with $r_z$ being the comoving distance of the light source away from us. This phenomena is known as ``cosmological birefringence\". This rotation angle $\\Delta\\alpha$ can be obtained by observing polarized radiation from distant sources such as radio galaxies, quasars and CMB. The Stokes parameters $Q$ and $U$ of the CMB polarization can be decomposed into a gradient-like ($G$) and a curl-like ($C$) component \\cite{Kamionkowski:1996ks}. For the standard theory of CMB, the $TC$ and $GC$ cross-correlation power spectra vanish. With the existence of cosmological birefringence, the polarization vector of each photon is rotated by an angle $\\Delta\\alpha$, and one would observe nonzero $TC$ and $GC$ correlations, even if they are zero at the last scattering surface. Denoting the rotated quantities with a prime, one gets \\cite{Feng:2004mq,Lue:1998mq}: \\begin{eqnarray}\\label{modify} C_{l}^{'TC} &=& C_{l}^{TG}\\sin(2\\Delta\\alpha)~, \\nonumber\\\\ C_{l}^{'GC} &=& \\frac{1}{2}(C_{l}^{GG}-C_{l}^{CC})\\sin(4\\Delta\\alpha)~,\\nonumber\\\\ C_{l}^{'TG} &=& C_{l}^{TG}\\cos(2\\Delta\\alpha)~,\\nonumber\\\\ C_{l}^{'GG} &=& C_{l}^{GG}\\cos^2(2\\Delta\\alpha) + C_{l}^{CC}\\sin^2(2\\Delta\\alpha)~,\\nonumber\\\\ C_{l}^{'CC} &=& C_{l}^{CC}\\cos^2(2\\Delta\\alpha) + C_{l}^{GG}\\sin^2(2\\Delta\\alpha)~, \\end{eqnarray} while the temperature power spectrum $TT$ remains unchanged. ", "conclusions": "" }, "0710/0710.3113_arXiv.txt": { "abstract": "{In April $2006$ a $4$-channel acoustic antenna has been put in long-term operation on Lake Baikal. The detector was installed at a depth of about $100$ m on the instrumentation string of Baikal Neutrino Telescope NT200+. This detector may be regarded as a prototype of subunit for a future underwater acoustic neutrino telescope. We describe the design of acoustic detector and present first results obtained from data analysis.} \\begin{document} ", "introduction": "\\begin{figure*}[t] \\begin{center} \\includegraphics [width=0.98\\textwidth]{icrc0639_fig01.eps} \\end{center} \\caption{Schematic view of underwater 4-channel digital device for detection of acoustic signals from high energy neutrinos.} \\label{fig1} \\end{figure*} The large scale neutrino telescopes currently under operation (NT200+ in Lake Baikal, AMANDA/IceCube at the South Pole and ANTARES in the Mediterranean) detect the Cherenkov light emitted in water or ice by relativistic charged particles produced via neutrino interactions with matter. Back in 1957, G.A. Askaryan has shown that a high-energy particle cascade in water should also produce an acoustic signal \\cite{Askarian-1957}. The absorption length for acoustic waves with a frequency about 30 kHz (the peak frequency of acoustic signals from a shower) in sea water is at least an order of magnitude larger than that of Cherenkov radiation, in the fresh Baikal water this ratio is even close to 100 \\cite{Clay}. Therefore acoustic pulses can be detected from considerably larger distances than Cherenkov radiation, and the acoustic method appears to be attractive for the detection of ultra high-energy neutrinos \\cite{Askarian-1977}. However, the technology of acoustic detection in high-energy physics is much worse developed than optical methods. Since several years, however, an increasing number of feasibility studies on acoustic particle detection are performed \\cite{ARENA}. In order to test the possibility of acoustic detection of high-energy neutrinos in Lake Baikal, the Baikal collaboration started with an in-situ study of acoustic noise which constitutes the background for the acoustic neutrino detection in the lake. For the purpose of noise measurement, an autonomous hydro-acoustic recorder with two input channels has been developed. We have performed a series of hydro-acoustic measurements in Lake Baikal in order to investigate the the background properties \\cite{Zeuthen-1, Akustika-noise}. It turned out that at stationary and homogeneous meteorological conditions the integral noise power in the frequency range $20$-$50$ kHz can reach levels as low as about $1$ mPa. At the same time, short acoustic pulses with different amplitudes and shapes including bipolar ones have been observed. The latter should be considered as a background for acoustic neutrino detection. However, the overwhelming majority of the short pulses have probably been generated by quasi-local sources or are due to interference of noise sound waves coming from a layer near the surface. Taking into account these properties of the noise, we conclude that the most promising way to detect acoustic particle signals is to deploy a net of rather compact acoustic antennas at relatively shallow depths (for example about $100$--$200$ m for Lake Baikal) and monitoring the water volume top-down. It is also necessary to suppress signals from the surface by caps made of a sound-absorbing material and mounted on top of the antennas. ", "conclusions": "The results of the experiment have demonstrated the feasibility of the proposed acoustic pulse detection technique in searching signals from cascade showers. Although the Baikal water temperature is close to the temperature of its maximal density, the absence of strong acoustic noise sources in the lake's deep zone, and the very low absorption of sound in freshwater may result in neutrino detection in Lake Baikal with a threshold as low as $10^{18}$ -- $10^{19}$ eV. This motivates further activities towards a large-scale acoustic neutrino detector in Lake Baikal." }, "0710/0710.5503_arXiv.txt": { "abstract": "We present relations of the black hole mass and the optical luminosity with the velocity dispersion and the luminosity of the \\nev\\ and the \\oiv\\ high-ionization lines in the mid-infrared (MIR) for 28 reverberation-mapped active galactic nuclei. We used high-resolution \\spi\\ Infrared Spectrograph and {\\it Infrared Space Observatory} Short Wavelength Spectrometer data to fit the profiles of these MIR emission lines that originate from the narrow-line region of the nucleus. We find that the lines are often resolved and that the velocity dispersion of \\nev\\ and \\oiv\\ follows a relation similar to that between the black hole mass and the bulge stellar velocity dispersion found for local galaxies. The luminosity of the \\nev\\ and the \\oiv\\ lines in these sources is correlated with that of the optical 5100$\\ang$ continuum and with the black hole mass. Our results provide a means to derive black hole properties in various types of active galactic nuclei, including highly obscured systems. ", "introduction": "\\label{sec:intro} Following the relation between the black hole (BH) mass, \\mbh, and the bulge stellar velocity dispersion, $\\sigma_*$, (\\citealt{ferrarese}; \\citealt{gebhardt}; \\citealt{tremaine}), a similar relation was found for the velocity dispersion $\\sigma$ of the \\oiii\\ 5007\\ang\\ line (\\citealt{nelson00}; \\citealt{greene}) that originates from the narrow-line region (NLR) gas of the active galactic nucleus (AGN). The NLR gas is thought to be mostly gravitationally bound to the bulge in high-luminosity AGNs (\\citealt{nelson96}; \\citealt{greene}; \\citealt{laor07}), unlike the broad-line region gas that is virialized due its proximity to the BH (\\citealt{peterson99}). This relation is useful for systems in which the stellar absorption lines cannot be observed, e.g., because of dilution of the stellar light by the AGN continuum. In addition to this relation, \\mbh\\ and the luminosity $L$ of the 5100 \\ang\\ optical continuum, \\lopt , are correlated in AGNs with BH measurements (\\citealt{kaspi00}). There are two main advantages to expanding such relations in the mid-infrared (MIR). The first is that the \\ion{Ne}{5} and \\ion{O}{4} ions emitting at 14.32 and 25.89 \\micron\\ cannot be easily excited by star-forming regions since they have ionization potentials $\\chi$ of 97.12 and 54.93 eV. The second reason is the low obscuration, which allows for the results to be applied to type 2 AGNs. In this Letter, we investigate for relations between \\mbh\\ and \\lopt\\ with MIR NLR line velocity dispersions and luminosities. ", "conclusions": "\\label{sec:origin} The relations between \\snev,\\soiv\\ and \\mbh\\ indicate that the NLR gas kinematics are primarily determined by the potential of the bulge, as it is also believed based on the \\oiii\\ line profiles (\\citealt{whittle92}; \\citealt{nelson96}; \\citealt{greene}). However, \\snev\\ and \\soiv\\ are on average 67 and 62 \\kms\\ higher than $\\sigma_*$, and 51 and 32 \\kms\\ higher than \\soiii . Such discrepancies could be attributed to an increase of the line width with increasing $\\chi$, which is often found for optical NLR lines (e.g. \\citealt{osterbrock}; \\citealt{whittle85b}). It is possible that ions with high ionization potentials have high velocity dispersions because they are located in NLR clouds that are close to the BH sphere of influence. Since the \\ion{O}{4} and, mostly, the \\ion{Ne}{5} ions are predominantly excited by AGNs, the MIR \\msigma\\ relation can be applied even in systems that undergo starbursts. It can also be applied in type 2 AGNs since the lines suffer little from extinction. Moreover, \\snev\\ and \\soiv\\ can be used as surrogates for $\\sigma_*$ in environments where measuring $\\sigma_*$ is hard, such as in bright QSOs. However, the relation does not necessarily hold for AGNs with strong winds and jets such as luminous radio sources (\\citealt{whittle92}), and for AGNs with high Eddington rates, \\eedd\\ (\\citealt{greene}). Such systems can be recognized by the asymmetric wings or the high kurtosis of their line profiles (Whittle 1985a; 1992). The luminosities of the broad lines (\\citealt{baldwin}) and the narrow lines (Fig.~\\ref{fig:ff}) scale with \\lopt\\ and with the bolometric AGN luminosity, \\lbol , since they depend on the absolute accretion rate onto the central object. Specifically, \\lbol\\ is equal to \\xopt \\lopt, \\xnev \\lnev, and \\xoiv \\loiv , where $x$ is a wavelength-dependent bolometric correction factor. For \\xopt $=$9 (\\citealt{kaspi00}; \\citealt{marconi04}) and for the median values of \\lopt/\\lnev\\ and \\lopt/\\loiv\\ in our sample, we find that \\xnev $=$13000 and \\xoiv $=$4500. That the relation does not have a slope of unity could indicate that \\xnev\\ and \\xoiv\\ depend on $L$ in a manner different than \\xopt\\ does. \\cite{netzer06} found that the equivalent width of \\oiii\\ decreases with increasing \\lopt. This behaviour could be attributed to a different geometric distribution of NLR clouds around the AGN at different luminosities. Whether the \\nev\\ or \\oiv\\ luminosity of a source can be used to derive its \\mbh\\ partly depends on its Eddington rate. To illustrate how changes of \\eedd\\ affect the relation between the luminosity of a MIR line and \\mbh , we overplot lines of constant \\eedd\\ in Figure~\\ref{fig:ml}. The more quiescent a source is, the more its position is shifted to the top left corner of this diagram. Equation~(\\ref{rel3}) is valid within the \\eedd\\ range 0.003$-$0.6 that our sources span, and implies that the mass of a BH determines its luminosity output. Its advantage is that it can be applied to sources with obscured optical continua or underestimated \\oiii\\ luminosities, such as type 2 AGNs (\\citealt{netzer06}). However its scatter is likely to be larger than that presented in \\S~\\ref{sec:mir_opt} since the Eddington rates of the reverberation-mapped AGNs are not necessarily representative of those of all local AGNs. We conclude that the MIR NLR gas kinematics trace \\mbh\\ in local reverberation-mapped AGNs in a manner similar to stellar kinematics. The \\nev\\ and \\oiv\\ line widths can be used to estimate \\mbh. The calibration of \\lnev\\ and \\loiv\\ to \\lopt\\ provides a new method to compute bolometric luminosities, black hole masses, and Eddington rates in various types of AGNs, including highly obscured systems, albeit with a large scatter.\\\\ \\\\ This work was based on observations made with the {\\it Spitzer Space Telescope}, % and was supported by NASA through an award issued by JPL/Caltech." }, "0710/0710.2115_arXiv.txt": { "abstract": "With a goal toward deriving the physical conditions in external galaxies, we present a survey of the formaldehyde emission in a sample of starburst systems. By extending a technique used to derive the spatial density in star formation regions in our own Galaxy, we show how the relative intensity of the $1_{10}-1_{11}$ and $2_{11}-2_{12}$ K-doublet transitions of H$_2$CO can provide an accurate densitometer for the active star formation environments found in starburst galaxies. Relying upon an assumed kinetic temperature and co-spatial emission and absorption from both H$_2$CO transitions, our technique is applied to a sample of nineteen IR-bright galaxies which exhibit various forms of starburst activity. In the five galaxies of our sample where both H$_2$CO transitions were detected we have derived spatial densities. We also use H$_2$CO to estimate the dense gas mass in our starburst galaxy sample, finding similar mass estimates for the dense gas forming stars in these objects as derived using other dense gas tracers. A related trend can be seen when one compares $L_{IR}$ to our derived $n(H_2)$ for the five galaxies within which we have derived spatial densities. Even though our number statistics are small, there appears to be a trend toward higher spatial density for galaxies with higher infrared luminosity. This is likely another representation of the $L_{IR}$-$M_{dense}$ correlation. ", "introduction": "\\label{intro} Studies of the distribution of Carbon Monoxide (CO) emission in external galaxies (\\cf\\ \\cite{Young1991}) have pointed to the presence of large quantities of molecular material in these systems. These studies have yielded a detailed picture of the molecular mass in many external galaxies. But, because emission from the abundant CO molecule is generally dominated by radiative transfer effects, such as high optical depth, it is not a reliable monitor of the physical conditions, such as spatial density and kinetic temperature, quantities necessary to assess the possibility of star formation. Emission from less-abundant, higher-dipole moment molecules are better-suited to the task of deriving the spatial density and kinetic temperature of the dense gas in our and external galaxies. For this reason, emission line studies from a variety of molecules have been made toward mainly nearby galaxies (see \\cite{Mauersberger1989} (CS), \\cite{Gao2004a} (HCN), \\cite{Nguyen1992} (HCO$^+$), \\cite{Mauersberger1990} and \\cite{Meier2005} (HC$_3$N), \\cite{Mauersberger2003} (NH$_3$), or \\cite{Henkel1991} for a review). The most extensive sets of measurements of molecular line emission in external galaxies has been done using the J=1-0 transitions of CO \\citep{Helfer2003} and HCN \\citep{Gao2004a}. Since the J=1-0 transitions of CO and HCN are good tracers of the more generally distributed and the denser gas, respectively, but do not provide comprehensive information about the individual physical conditions of the dense, potentially star-forming gas, another molecule must be observed for this purpose. Formaldehyde (H$_2$CO) has proven to be a reliable density and kinetic temperature probe in Galactic molecular clouds. Existing measurements of the H$_2$CO $1_{10}-1_{11}$ and $2_{11}-2_{12}$ emission in a wide variety of galaxies by \\cite{Baan1986}, \\cite{Baan1990}, \\cite{Baan1993}, and \\cite{Araya2004} have mainly concentrated on measurements of the $1_{10}-1_{11}$ transition. One of our goals with the present study was to obtain a uniform set of measurements of both K-doublet transitions with which the physical conditions, specifically the spatial density, in the extragalactic context could be derived. Using the unique density selectivity of the K-doublet transitions of H$_2$CO we have measured the spatial density in a sample of galaxies exhibiting starburst phenomena and/or high infrared luminosity. In \\S\\ref{H2coProbe} we discuss the specific properties of the H$_2$CO molecule which make it a good probe of spatial density. \\S\\ref{Observations} presents our observation summary; \\S\\ref{Results} our H$_2$CO, OH, H111$\\alpha$, and continuum emission measurement results; \\S\\ref{Analysis} analyses of our H$_2$CO, OH, and H111$\\alpha$ measurements, including Large Velocity Gradient (LVG) model fits to and dense gas mass calculations based on our H$_2$CO measurements. ", "conclusions": "\\label{Conclusions} Using measurements of the $1_{10}-1_{11}$ and $2_{11}-2_{12}$ K-doublet transitions of H$_2$CO we have derived accurate \\textit{measurements} of the spatial density (n(H$_2$)) in a sample of starburst galaxies. The derived densities range from $10^{4.7}$-$10^{5.7}$~cm$^{-3}$, consistent with the suggestion that the high infrared brightness of these galaxies is driven by extreme star formation activity. We believe that these spatial density measurements are the most accurate measurements of this important physical quantity in starburst galaxies made to-date. We have also used our H$_2$CO measurements to derive a measure of the dense gas mass which ranges from $0.06-77\\times10^8 M_\\odot$, generally consistent with previous measurements, mainly from studies of the HCN emission in these galaxies. The linear correlation between the IR luminosity ($L_{IR}$) and the dense gas mass ($M_{dense}$) found in larger samples of the HCN emission in starburst galaxies is also apparent in our H$_2$CO measurements. This further supports the suggestion that active star formation in IR-bright galaxies is driven by the amount of material available to form stars. We also note a related trend between $L_{IR}$ and our derived $n(H_2)$ for the five galaxies within which we have derived spatial densities. The three galaxies with lower IR luminosities have lower ($n(H_2) \\simeq 10^5$~cm$^{-3}$) derived spatial densities, while the two higher luminosity galaxies have higher ($n(H_2) = 10^{5.7}$~cm$^{-3}$) derived spatial densities. This is likely another representation of the $L_{IR}$-$M_{dense}$ correlation." }, "0710/0710.0867_arXiv.txt": { "abstract": "The observed cosmic acceleration presents the physics and cosmology communities with amazing opportunities to make exciting, probably even radical advances in these fields. This topic is highly data driven and many of our opportunities depend on us undertaking an ambitious observational program. Here I outline the case for such a program based on both the exciting science related to the cosmic acceleration and the impressive impact that a strong observational program would have. Along the way, I challenge a number of arguments that skeptics use to question the value of a strong observational commitment to this field. ", "introduction": "This is truly remarkable time to be involved in cosmology research. There are many reasons for this, but none stand out quite as dramatically as the observed cosmic acceleration (attributed in current nomenclature to the ``dark energy''). In the words of the Dark Energy Task Force (DETF)~\\cite{Albrecht:2006um} ``most experts believe that nothing short of a revolution in our understanding of fundamental physics will be required to achieve a full understanding of the cosmic acceleration''. As things stand, this revolution is being motivated both by remarkable new data sets and by exciting theoretical developments. Many ambitious researchers in cosmology and related fields have been galvanized by these extraordinary developments. There have been a number of excellent talks at this conference on the topic of cosmic acceleration which capture some of this climate. The DETF, charged with charting a way forward with dark energy observations, received 50 thoughtful and thorough whitepapers from leaders in the field (despite the lack of any specific commitment at that time to fund future dark energy experiments). Most of us look at these developments and are astonished at our good fortune to be part of what will surely be viewed as one of the great moments in the history of science. Indeed, since the discovery of dark energy every group that has deliberated on future directions for the field has recognized the exciting opportunities and challenges presented by the cosmic acceleration\\footnote{See for example \\cite{Turner:2003pe,Albrecht:2005np,Peacock:2006kj}. Since the 2006 DETF report a number of panels have recommended pursuit of specific ground and space based dark energy projects, and just since PASCOS 07 the US National Research Council's {\\em Committee on NASA's Beyond Einstein Program} named a ``Joint Dark Energy Mission'' the top priority for that program \\url{http://nationalacademies.org/morenews/20070907b}}. But a small number of negative voices continue to be heard alongside the building enthusiasm for further studies of the dark energy. To some degree this negativity may reflect the natural skepticism of scientists in the face of unbridled enthusiasm (such as the above). To the extent that that is the explanation, the skepticism is surely a good thing that will lead to a healthy debate and produce more rigorous research. On the other hand, I suspect some of the negativity is just the result of sloppy thinking and needs to be challenged and simply put to rest. Regardless of how one might attribute explanations and motivations, one purpose of this paper is to engage in this debate on a number of fronts. Here are some illustrations of some of the negative views I am talking about. At lunch on the first day of the PASCOS 07 conference there was a lively discussion about the merits of dark energy experiments. One colleague commented \\begin{quote} {\\em ``Studies of dark energy are unlikely to be interesting because we already have a theory of dark energy.''} \\end{quote} moments later, another cosmologist declared \\begin{quote} { \\em ``Studies of dark energy are unlikely to be interesting because we have no theory of dark energy.''} \\end{quote} The two were united in their conclusion, and apparently not too bothered by subtle differences in their reasoning One unavoidable feature of physics research is that at any given time the total amount of data is finite. That necessarily means there will be more than one theory that fits the data. In many fields, this universal fact is seen as a source of vitality. Curiosity about further resolving these degeneracies drives exciting new experiments and theoretical work. The momentous progress in physics over the last century can be seen in this light as the fundamental particles, the atomic theory of heat, quantum physics and general relativity emerged from ``under the radar'' of earlier data sets that were fit perfectly by more primitive theories. As was amply evident in the talks at the PASCOS 07 conference, we have good reason to look forward to similar advances as the LHC data starts coming in. But for some reason, the fact that a given future dark energy experiment will not remove all uncertainties about the nature of dark energy seems to generate considerable angst among some physicists and astronomers\\footnote{See the question section of my PASCOS 07 talk for an example\\cite{PASCOS07}} and is sometimes given as a reason to be discouraged from even doing more experiments. As I shall quantify below (and as was also shown by others at PASCOS 07), the proposed new experiments will have an impressive impact on our knowledge of dark energy. This is all one can ever ask of a new experiment. This paper focuses on two key areas. In the next section I review some of the thriving theoretical work that has been stimulated by the cosmic acceleration. The striking theoretical issues raised by the cosmic acceleration and the remarkable directions we have been driven in our initial attempts to understand it are the key reasons I find this topic so deeply interesting. I also believe this is why so many excellent researchers are taking risks and changing direction in their careers in order to get involved. The third section summarizes a number of results which demonstrate the tremendous impact future experiments can have on our understanding of dark energy, both on our understanding of the general properties of dark energy and in terms of constraining specific models of dark energy that are currently of interest. It is these results that demonstrate that an ``aggressive program of dark energy probes'' is indeed possible. This paper is based on my talk at the PASCOS 07 meeting at Imperial College. My slides and a video of the talk are available online\\cite{PASCOS07}. This online material as well as my ``Origins of Dark Energy'' talk\\cite{ODE} and related papers\\cite{Albrecht:2007qy,Abrahamse:2007ip} are a good source for the technical material on which this paper is based. My goal here is to assemble some key arguments in a concise form. Readers seeking more details should refer to this other material. ", "conclusions": "\\label{Sect:CAS} The case for aggressive pursuit of new data on dark energy is twofold. Firstly, I have outlined how the subject of dark energy has generated very exciting and often radical new theoretical ideas. These include dramatic proposals that change how we think about equilibrium and initial conditions in cosmology and even how we formulate fundamental theories. Second, impressive new experiments are within reach that could have a tremendous impact on our understanding of dark energy. The best experiments will constrain dark energy properties orders of magnitude better than the current or medium sized future experiments, and will have the ability to strongly discriminate among and even fully eliminate popular dark energy models based on subtle variations in the equation of state. I have outlined and challenged some of the skeptical perspectives I have heard regarding future dark energy studies. The discovery of the cosmic acceleration has caused a great upheaval in our thinking about fundamental physics and cosmology. I often sense that the skeptics are hoping this upheaval will end quickly and are grasping for arguments that will allow things to rapidly return to normal. I feel this outcome is very unlikely, and this is exactly why I find the topic so exciting. Nature has handed us an amazing opportunity. I hope that the physics and cosmology communities have the strength to face the challenge of the cosmic acceleration head on and give a response that we can be proud of when people write the history of this era. \\begin{theacknowledgments} I would like to thank A. Abrahamse, M. Barnard, G. Bernstein, B. Bozek, L. Sorbo, M. Yashar and the DETF members who collaborated with me on some of the work reviewed here, and B. Bozek for helpful comments on the manuscript. Also, I thank the organizers, especially Arttu Rajantie, for a really excellent conference. This work was supported in part by DOE grant DE-FG03-91ER40674 and NSF grant AST-0632901. \\end{theacknowledgments}" }, "0710/0710.5359_arXiv.txt": { "abstract": "{% The study of rotation and activity in low-mass stars or brown dwarfs of spectral classes M and L has seen enormous progress during the last years. I summarize the results from different works that measured activity, rotation, and sometimes magnetic fields. The generation of magnetic activity seems to be unchanged at the threshold to completely convective stars, i.e. no change in the efficiency of the magnetic dynamos is observed. On the other hand, a sudden change in the strength of rotational braking appears at the threshold mass to full convection, and strong evidence exists for rotational braking weakening with lower mass. A probable explanation is that the field topology changes from dipolar to small scale structure as the objects become fully convective. } ", "introduction": "Rotation and activity are intimately connected in sun-like stars. Rotation is believed to generate a magnetic field through a dynamo mechanism that scales with rotation. The stellar wind couples to the rotating magnetic field lines carrying away angular momentum so that the star is being braked. Hence stars of spectral type F--K rotate more rapidly and are more active when they are young, but they decelerate and become less active as they age; this is the so-called rotation-activity connection (Noyes et al., 1984; Pizzolato et al., 2003). The scaling of activity with rotation depends on the type of dynamo inside the star. The fact that the rotation-activity connection is similar in virtually all stars that harbor convective envelopes and at all ages (Pizzolato et al., 2003; Reiners, 2007) indicate that the dynamo mechanism is comparable in all these stars. Around spectral type M3.5, the internal structure of the stars changes; stars later than around M3.5 are believed to be fully convective. They cannot harbor an interface dynamo working at the tachocline, which is believed to be the most important dynamo mechanism in the hotter sun-like stars. If the interface dynamo was the only important mechanism driving a magnetic dynamo, one could expect a sharp break in magnetic field generation around spectral type M3.5. Such a break would imply a sudden change in observable stellar activity and in the braking of stellar rotation. A break in stellar activity is not observed in X-rays or H$\\alpha$ in the surveys that crossed the M3.5 border (e.g., Delfosse et al., 1998; Mohanty \\& Basri, 2003; West et al., 2004). Activity rather stays at comparable levels if normalized to bolometric luminosity, and quite surprisingly the fraction of active stars even raises up to $\\sim 80\\,\\%$ in late M-type objects before it goes down again. This clearly shows that the interface dynamo is not the only dynamo operating in stellar interiors and that fully convective stars can have quite efficient dynamos as well. On the other hand, Delfosse et al., 1998, presented a plot that shows a different behavior in rotational braking in stars later than spectral type M3.5. In their Fig.\\,3, they show that all M dwarfs earlier than M3.5 are slow rotators ($v\\,\\sin{i} \\la 3$\\,km\\,s$^{-1}$) regardless of what disk population they belong to (young or old). M stars later than M3.5, however, show substantial rotation velocities of up to $v\\,\\sin{i} = 50$\\,km\\,s$^{-1}$ in the young disk population, and in the old disk population some late M dwarfs still have velocities around $v\\,\\sin{i} = 10$\\,km\\,s$^{-1}$. This can be interpreted as an indication for a sudden change in the timescales of rotational braking at the mass where stars become fully convective. Delfosse et al. conclude that spin-down timescales are on the order of a few Gyrs at spectral type M3--M4, and of the order of 10\\,Gyr at spectral type M6. The investigation of rotation and activity in ultra-cool stars (M7 and later) was put on firm ground by Mohanty \\& Basri, 2003. From high-resolution spectra they determined projected rotation velocities, and from H$\\alpha$ emission they derived the level of activity. Reiners \\& Basri, 2007, added more M stars to this sample. In addition, they measured magnetic fields in low-mass M dwarfs and showed that in M dwarfs the level of activity is still coupled to magnetic flux -- high magnetic flux levels lead to strong H$\\alpha$ emission and stars without H$\\alpha$ emission show no magnetic fields. ", "conclusions": "High resolution spectroscopy in M and L dwarfs becomes a technique that allows to investigate the physics and the evolution of low mass objects in great detail. No change in H$\\alpha$ activity is observed at the threshold to complete convection. Activity can be followed to objects as late as mid-M and it might continue to even lower masses but below the current observational threshold. The decline in normalized activity among the ultra-cool dwarfs could be explained by the enhanced electrical resistivity at the low temperatures. Currently, no indications for a mass or temperature dependence of stellar or substellar dynamos can be concluded from the activity measurements. A sudden change in the behavior of rational braking at spectral class M3.5 was already found by Delfosse et al., 1998. In the young disk, stars earlier than M3.5 rotate slowly while later stars still show significant rotation. The lack of slowly rotating ultra-cool dwarfs and the rise of the minimum rotation rate with spectral class could be explained by rotational braking that is weaker with lower mass or lower temperature. The reason for a weaker braking at later spectral type could be a different magnetic topology. The geometry of the magnetic field is essential for the strength of magnetic braking as described in Krishnamurti et al. (1997) and Sills et al. (2000). These authors find that the description of magnetic braking ($\\omega_\\mathrm{crit}$) needs to be different in mid-M stars and in earlier objects, which they connect to different convective turnover times. Although the details of magnetic braking in ultra-cool dwarfs are not quite understood, a substantial amount of evidence exists that rotational braking is weaker with lower mass or lower temperature. The limiting factor of rotational braking may be the topology of the magnetic fields which in fully convective stars might be generated on smaller scales so that the topology is different from a dipolar configuration leading to the weaker rotational braking." }, "0710/0710.3396_arXiv.txt": { "abstract": "We modify a stellar structure code to estimate the effect upon the main sequence of the accretion of weakly interacting dark matter onto stars and its subsequent annihilation. The effect upon the stars depends upon whether the energy generation rate from dark matter annihilation is large enough to shut off the nuclear burning in the star. Main sequence WIMP burners look much like protostars moving on the Hayashi track, although they are in principle completely stable. We make some brief comments about where such stars could be found, how they might be observed and more detailed simulations which are currently in progress. Finally we comment on whether or not it is possible to link the paradoxically young OB stars found at the galactic centre with WIMP burners. ", "introduction": " ", "conclusions": "" }, "0710/0710.3169_arXiv.txt": { "abstract": "We compute the electromagnetic radiative corrections to all leading annihilation processes which may occur in the Galactic dark matter halo, for dark matter in the framework of supersymmetric extensions of the Standard Model (MSSM and mSUGRA), and present the results of scans over the parameter space that is consistent with present observational bounds on the dark matter density of the Universe. Although these processes have previously been considered in some special cases by various authors, our new general analysis shows novel interesting results with large corrections that may be of importance, e.g., for searches at the soon to be launched GLAST gamma-ray space telescope. In particular, it is pointed out that regions of parameter space where there is a near degeneracy between the dark matter neutralino and the tau sleptons, radiative corrections may boost the gamma-ray yield by up to three or four orders of magnitude, even for neutralino masses considerably below the TeV scale, and will enhance the very characteristic signature of dark matter annihilations, namely a sharp step at the mass of the dark matter particle. Since this is a particularly interesting region for more constrained mSUGRA models of supersymmetry, we use an extensive scan over this parameter space to verify the significance of our findings. We also re-visit the direct annihilation of neutralinos into photons and point out that, for a considerable part of the parameter space, internal bremsstrahlung is more important for indirect dark matter searches than line signals. ", "introduction": "During the last few years a strong consensus has emerged about the existence of a sizeable dark matter contribution to the total cosmological energy density. The identification of experimental signatures that eventually may determine the nature of the cosmological dark matter is thus becoming ever more important. The present estimates \\cite{wmap} give the fraction of the critical density of cold dark matter particles as $\\Omega_{CDM}h^2\\sim 0.105 \\pm 0.013$, where the Hubble parameter (scaled in units of 100 km/s Mpc$^{-1}$) is $h\\sim 0.70\\pm 0.02$. Also, on the scales of galaxies and smaller, a number of methods including measurements of rotation curves as well as gravitational lensing agree well with the predictions from N-body calculations of gravitational clustering in cold dark matter cosmologies (see e.g. \\cite{vialactea}). The methods of detection of dark matter (for reviews, see \\cite{reviews}) can be divided into {\\em accelerator} production and detection of missing energy (especially at the LHC at CERN, which will start operating some time in 2008), {\\em direct detection} (of dark matter particles impinging on a terrestrial detector, with recent impressive upper limits reported by \\cite{direct}), or {\\em indirect detection} of particles generated by the annihilation of dark matter particles in the Galactic halo or in the Sun/Earth. All these methods are indeed complementary -- it is probable that a signal from more than one type of experiment will be needed to fully identify the particle making up the dark matter. The field is just entering very interesting times, with the LHC soon starting and new detectors of liquid noble gases being developed for direct detection. For indirect detection, the satellite PAMELA \\cite{pamela} was launched a year ago and will soon reveal its first sets of data for positron and antiproton yields in the cosmic rays \\cite{antimatter}. AMANDA \\cite{amanda} at the South Pole that has searched for detection of neutrinos from the centre of the Earth or the Sun \\cite{neutrinos}, will soon give way to the much larger detector IceCUBE \\cite{icecube}, and for gamma-rays coming from annihilations of dark matter particles in the halo \\cite{gammas} the space satellite GLAST \\cite{glast}, to be launched in 2008, will open up a new window to the high-energy universe, for energies from below a GeV to about 300 GeV. One problem with all these discovery methods is that the signal searched for may be quite weak, with much larger backgrounds in many cases. For indirect detection through gamma-rays, the situation may in principle be better, due to (i) the direct propagation from the region of production, without significant absorption or scattering; (ii) the dependence of the annihilation rate on the square of the dark matter density which may give \"hot spots\" near density concentrations as those predicted by N-body simulations; (iii) possible characteristic features like gamma-ray lines or steps, given by the fact that no more energy than $m_{\\chi}$ per particle can be released in the annihilation of two non-relativistic dark matter particles (we denote the dark matter particle by $\\chi$). As an example, it was recently shown \\cite{idm} that in models of an extended Higgs sector, the line signal from the two-body final states $\\gamma\\gamma$ and $Z\\gamma$ could give a spectacular signature in the gamma-ray spectrum between 40 and 80 GeV. On the other hand, in models of universal extra dimensions (UED) \\cite{uedline} or in the theoretically perhaps most favoured, supersymmetric, models of dark matter the line feature is in general not very prominent, except in some particular regions of the large parameter space. However, it was early realised that there could be other important spectral features \\cite{lbe89}, and recently it has been shown that internal bremsstrahlung (IB) from produced charged particles in the annihilations could yield a detectable \"bump\" near the highest energy for heavy gauginos or Higgsinos annihilating into $W$ boson pairs, such as expected in split supersymmetry models \\cite{heavysusy}. In \\cite{birkedal}, it was furthermore pointed out that IB often can be estimated by simple, universal formulas and often gives rise to a very prominent step in the spectrum at photon energies of $E_\\gamma=m_\\chi$ (such as in UED models \\cite{Bergstrom:2004cy}). Encouraged by these partial results, we have performed a detailed analysis of the importance of IB in the minimal supersymmetric extension to the standard model (MSSM). We have therefore calculated the IB contributions for all two-particle charged final states from general neutralino annihilations. Besides confirming the mentioned partial results for the universal radiative corrections, in particular those relating to soft and collinear bremsstrahlung, we also point out interesting cases of model-dependent ``virtual'' brems\\-strahlung (i.e. photons emitted from charged virtual particles), see Fig.~1. We confirm the suspicion expressed already in \\cite{lbe89} that this type of emission may circumvent the chiral suppression, i.e., the annihilation rate being proportional to $m_f^2$ for annihilation into a fermion pair from an $S$-wave initial state, as is the case in lowest order for non-relativistic dark matter Majorana particles in the Galactic halo (see also \\cite{baltz_bergstrom}). Since this enhancement mechanism is most prominent in cases where the neutralino is close to degenerate with charged sleptons, it is of special importance in the so-called stau coannihilation region in models of minimal supergravity (mSUGRA, as implemented in \\cite{isajet}). We therefore run through an extensive scan over these models (based on \\cite{baltz_peskin}) and find, indeed, remarkable cases of enhancement of the gamma-ray rate in the stau coannihilation region, near the maximal possible photon energy $E_\\gamma=m_\\chi$. Let us stress that the radiative corrections to the main annihilation channels, here computed systematically for the first time, may turn out to be of utmost importance when fitting gamma-ray data, e.g. from GLAST, to supersymmetric dark matter templates. Over much of the parameter space we have scanned, these corrections give a large factor of enhancement over the commonly adopted estimates, especially at the observationally most interesting, highest energies. More importantly, they add a feature, the very sharp step at the dark matter mass, that would distinguish this signal from all other astrophysical background (or foreground) processes. ", "conclusions": "As can be seen already from our benchmark points in Table~\\ref{benchmark}, and in more detail from the scatter plots in Fig.~\\ref{FSRcomp1}, the internal bremsstrahlung effects computed in this work can be very significant, changing sometimes by more than an order of magnitude the lowest-order prediction for the high-energy gamma-ray signal from neutralino dark matter annihilation. Although some of these enhancements have been found before \\cite{lbe89,heavysusy,birkedal}, this is the first time the first-order radiative corrections have been computed systematically, for all relevant final states in supersymmetric dark matter models. The resulting enhancements of the expected fluxes are surprisingly large over significant regions in the parameter space of the MSSM, including the more constrained mSUGRA models. Despite the fact that some large corrections apply to absolute rates that are too small to be of practical interest, Fig.~4 shows that the quantity ${\\cal S}$, which is directly proportional to the expected signal in gamma-ray detection experiments, also is significant for the internal bremsstrahlung contribution in large regions of parameter space. For $m_\\chi < 300$ GeV, for example, values of ${\\cal S}_{IB}$ greater than 0.1 are generic, and for masses below 100 GeV, values of 1 or higher are common, which in very many cases is higher than the corresponding values for the line signals $\\gamma\\gamma$ and $Z\\gamma$. One should also bear in mind that the sensitivity of Air Cherenkov Telescopes increases significantly with energy; detectional prospects for a $m_\\chi\\sim1~$TeV neutralino with ${\\cal S}\\sim0.01$, e.g., correspond very roughly to those for a $m_\\chi\\sim100~$GeV neutralino with ${\\cal S}\\sim0.5$ (see, e.g., \\cite{Morselli:2002nw}). In this light, the situation becomes very interesting even for TeV scale Higgsinos, where IB generically contributes more than 10 times as much as secondary photons. We note that (as anticipated in \\cite{lbe89}) helicity suppression and also CP selection rules of certain final states may be circumvented by emitting a photon; this is for example the origin of the very substantial enhancements of the signal obtained in the stau annihilation region in mSUGRA models. In this situation, the probability of emitting gamma rays vanishes at zero photon energy but increases rapidly at high energy (see Fig.~\\ref{fig:spec}b and \\ref{fig:spec}c), which gives a photon ``bump'' at $E_\\gamma\\approx m_\\chi$. We also note that in less constrained versions of the MSSM than considered here, we expect even more situations where large enhancements of the annihilation signal due to internal bremsstrahlung can be found. An example is heavy Wino dark matter \\cite{heavysusy,Chattopadhyay:2006xb}, which becomes possible when relaxing the condition $M_1\\approx \\frac{1}{2}M_2$ (as realized, e.g., in anomaly mediated supersymmetry breaking scenarios \\cite{amsmb}). Of course, the line signals, in particular $\\gamma\\gamma$, have the virtue of being at the highest possible energy, so in order to make a more accurate comparison between these and the IB signal computed here, one would have to model also the expected spectral shape of possible astrophysical gamma-ray backgrounds and the energy resolution of the detector. This is left for future work \\cite{torstenetal}. We note, however, that in general also the new contributions have a characteristic signature, ({\\em cf.} Fig.~2) which can hardly be mimicked by any known astrophysical gamma-ray source. In fact, in some cases these spectra could even be used by future experiments to distinguish between different dark matter candidates (note that, e.g., the distinction between Kaluza-Klein dark matter and a neutralino in the focus point region like our BM4 point would be possible already with the energy resolution of present Air Cherenkov Telescopes \\cite{Bergstrom:2006hk}). To conclude, we have shown that the commonly neglected first-order radiative corrections to neutralino dark matter annihilation should definitely be taken into account when predicting rates for gamma-ray telescopes. In particular, the soon to be launched GLAST space telescope \\cite{glast} will have an enhanced possibility over what has previously been assumed to detect radiation from supersymmetric dark matter annihilation. The routines needed to compute these new processes will be included in the next release of the \\ds\\ package \\cite{ds,joakim}. \\smallskip" }, "0710/0710.4922_arXiv.txt": { "abstract": "Using the high-resolution spectrometer SPI on board the \\emph{International Gamma-Ray Astrophysics Laboratory (INTEGRAL)}, we search for a spectral line produced by a dark matter (DM) particle with a mass in the range $40 keV < M_{DM} < 14 MeV$, decaying in the DM halo of the Milky Way. To distinguish the DM decay line from numerous instrumental lines found in the SPI background spectrum, we study the dependence of the intensity of the line signal on the offset of the SPI pointing from the direction toward the Galactic Centre. After a critical analysis of the uncertainties of the DM density profile in the inner Galaxy, we find that the intensity of the DM decay line should decrease by at least a factor of 3 when the offset from the Galactic Centre increases from $0^\\circ$ to $180^\\circ$. We find that such a pronounced variation of the line flux across the sky is not observed for any line, detected with a significance higher than $3\\sigma$ in the SPI background spectrum. Possible DM decay origin is not ruled out only for the unidentified spectral lines, having low ($\\sim 3\\sigma$) significance or coinciding in position with the instrumental ones. In the energy interval from 20 keV to 7 MeV, we derive restrictions on the DM decay line flux, implied by the (non-)detection of the DM decay line. For a particular DM candidate, the sterile neutrino of mass $M_{DM}$, we derive a bound on the mixing angle. ", "introduction": "\\subsubsection*{Dark matter in the Universe} There is a vast body of evidence, suggesting that the large fraction of matter in the Universe exists in the form of the \\emph{Dark matter} \\emph{(DM)}. However, while the total density of the DM is measured with a very high precision ($\\Omega_{\\rm DM}h^2 = 0.105^{+0.007}_{-0.009}$,~\\citealt{WMAP3}), little is known about its properties apart from this. The possibility that the DM is composed of the Standard Model (SM) particles has been ruled out for a long time already. Indeed, the DM cannot be made out of baryons, as producing such an amount of baryonic matter would require drastic modifications of the scenario of the Big Bang nucleosynthesis (BBN), which otherwise successfully describes the abundance of light elements~(see for example ~\\citealt{Dar:95}). Recent microlensing experiments rule out the possibility that another type of baryonic DM -- massive compact halo objects (MACHOs) -- constitute dominant fraction of mass in the halo~\\citep{MACHO:00,EROS:00,OGLE:98}. The only non-baryonic DM candidate in the SM candidates -- (left-handed) neutrino -- is ruled out from the large scale structure (LSS) considerations~\\citep[see e.g.][]{Bond:80,Hannestad:03,Crotty:04}. What are the properties of a successful DM candidate? First of all, this particle should be massive. Many extensions of the SM present the DM candidates with the masses ranging from $\\sim 10^{-10}\\ev$~(massive gravitons, \\citealt{Dubovsky:04}) and $\\sim 10^{-6}$~eV (axions) to hundreds of GeV (WIMPs) and even to $10^{13}$~GeV \\citep[WIMPZILLA,][]{Kuzmin:98,Kuzmin:99,Chung:98}. For a review of particle physics DM candidates see e.g.~\\cite{Bergstrom:00,Bertone:04,Carr:06}. Secondly, there should exist mechanisms of DM production with the correct abundances. The production mechanism in particular determines the velocity distribution of particles in the early Universe. This velocity distribution can, in principle, be probed experimentally. Namely, if during the structure formation epoch the DM particles have velocities, comparable to the speed of sound in the baryon-photon plasma, they ``erase'' density fluctuations at scales, smaller than the distance, they have traveled (called the \\emph{free-streaming length}). To differentiate various models in accordance with this property, the DM candidates with the negligible velocity dispersion (and, correspondingly, free-streaming) are called \\emph{cold} DM (CDM), while those with the free-streaming of the order of $\\sim 1\\mpc$ are considered to be \\emph{warm} (WDM).\\footnote{The left-handed neutrino would represent \\emph{hot} DM in this terminology, i.e. the DM with the free-streaming length $\\gg 1$~Mpc.} It is possible to constrain the free-streaming length of a particular DM candidate by probing the structure of the Universe at galaxy-size scales. This can be done through the analysis of the Lyman-$\\alpha$ forest data~\\citep{Hui:97}. Lyman-$\\alpha$ analysis puts an upper bound on the free-streaming of the DM particles~\\citep{Hansen:01,Viel:05,Seljak:06,Viel:06,Viel:07}. It should be noted however that currently existing interpretation of the Lyman-$\\alpha$ data is model-dependent, as, apart from a number of astrophysical assumptions (see~\\citealt{Hui:97}) and complicated hydrodynamic simulations, it relies on {\\it a priori} assumptions about the velocity distribution of the DM particles. A way to differentiate between CDM and WDM models would be to compare the numerical simulations of the DM distribution in the Milky Way-type galaxies with the actual observations. However, the resolution of the N-body simulations is not yet sufficient to answer the questions about e.g. the DM density profiles in dwarf satellite galaxies. Moreover, most of the simulations include only collisionless DM particles, and do not model the baryons and their feedback on the galaxy structure formation. These problems are not solved even for the CDM simulations, and WDM simulations have additional serious difficulties. From an observational point of view, it has been argued for some time already that there is a discrepancy between CDM simulations and observations (see e.g.~\\citealt{Moore:94,Moore:99,Klypin:99,Bode:00,Avila-Reese:01,Goerdt:06}) It has been claimed recently that a number of recent observations of dwarf satellite galaxies of the Milky way and Andromeda galaxy seem to indicate the existence of the smallest scale at which the DM exists~\\citep{Gilmore:06,Gilmore:07a,Gilmore:07b,Koposov:07}. However, this statement and the interpretation of the observations are still subject to debate~\\citep{Klimentowski:06,Penarrubia:07,Strigari:07,Simon:07}. Therefore it is too early to say what kind of DM models is favoured by comparing simulations and observations. Usually it is also necessary for the DM candidate to be stable. For the most popular DM candidate -- weakly interacting massive particles (WIMPs), this is related to the fact that the particles of $\\sim$ electroweak mass, having weak strength interaction with SM matter (required to produce the correct amount of DM), would decay too fast and would not be ``dark''. If, however, the DM particle interacts with the SM more weakly than WIMPs, it could well have a finite (although cosmologically long) life time. There exist several unstable (decaying) DM candidates e.g. gravitino~\\citep{Borgani:96,Baltz:01,Roszkowski:04,Cerdeno:05,Cembranos:06,Lola:07}. In this paper we will concentrate mainly on one candidate, the sterile neutrino (although our results will be applicable for any type of decaying DM). Constraints on the decaying DM were analyzed in~\\cite{DeRujula:80,Berezhiani:87,Doroshkevich:89,Berezhiani:90a,Berezhiani:90b,Bertone:07,Zhang:07} (see also the book by~\\citealt{Khlopov:97}). \\subsubsection*{Sterile neutrino DM} \\label{sec:sterile-neutrino-dm} It was noticed long ago that the right-handed (or as it is often called \\emph{sterile}) neutrino with the mass in the keV range would represent a viable DM candidate~\\citep{Dodelson:93}. Such a neutrino would interact with the rest of the matter only via the quadratic mixing with left-handed (\\emph{active}) neutrinos and therefore (although not stable) could have cosmologically long life-time. At the same time, it could be produced in the early Universe with the correct abundances~\\citep{Dodelson:93,Shi:98,Shaposhnikov:06}. One of the decay channels of the unstable sterile neutrinos includes emission of photons of the energy equal to half of the sterile neutrino rest energy. This potentially provides a possibility to observe the decays of DM sterile neutrinos via detection of a characteristic spectral line in the % spectra of astrophysical objects with large DM concentration. Recently this DM candidate has attracted much attention (see e.g.~\\citet{Shaposhnikov:07a} and references therein). It was found that a very modest and natural extension of the SM by 3 right-handed neutrinos (making the SM more symmetric as all SM fermions, including neutrino, would have now their left and right handed counterparts) provided a viable extension of the theory, capable of solving several ``beyond the SM'' problems. First of all, such an extension makes neutrinos massive and thus perhaps provides the simplest and the most natural explanation of the phenomenon of ``neutrino oscillations''~(see e.g.~\\cite{Fogli:05,Strumia:06,Giunti:06} for reviews). The smallness of neutrino masses in this model (called \\numsm in~\\citealt{Asaka:05b}) is achieved by the usual see-saw mechanism with Majorana masses of right-handed neutrinos being below electroweak scale.\\footnote{The fact that the \\numsm does not introduce any new scale above the electroweak one, makes this theory especially appealing from the point of view of its experimental verification/falsification.} Secondly, if two heavier sterile neutrinos ($N_2$ and $N_3$) are almost degenerate in mass and have their masses between $\\mathcal{O}(100)\\mev$ and $\\mathcal{O}(20)\\gev$, the \\numsm provides the mechanism of generating the baryon asymmetry of the Universe. Thirdly, the lightest sterile neutrino $N_1$ can have arbitrary mass and arbitrarily weak coupling with the (active) neutrino sector. At the same time, it can be produced in the early Universe in the correct amounts. It represents therefore the DM particle in the \\numsm. Thus, altogether the \\numsm represents (arguably) the simplest extension of the SM, capable of explaining three important questions: origin and smallness of neutrino masses, baryon asymmetry in the Universe and the existence of the DM. \\subsubsection*{Existing restrictions on sterile neutrino DM parameters.} What are the current restrictions on parameters (mass and \\emph{mixing}) of sterile neutrino DM? First of all sterile neutrino mass should satisfy the universal Tremaine-Gunn lower bound:\\footnote{In its simplest form the Tremaine-Gunn bound comes from the fact that for the fermions there is a maximal density in the phase space~\\citep{Tremaine:79,Dalcanton:00} and therefore the observed phase-space density in various DM dominated systems should be less that this (mass dependent) bound.} $\\mdm\\gtrsim 300-500\\ev$.\\footnote{A stronger lower bound from Ly-$\\alpha$~\\citep{Seljak:06,Viel:06,Viel:07} can be obtained in the case of the particular production mechanisms -- the Dodelson-Widrow scenario~\\citep{Dodelson:93}. For other possible production mechanisms~\\citep[e.g.][]{Shi:98,Shaposhnikov:06} the Ly-$\\alpha$ constraints should be reanalyzed.} Next, as the sterile neutrino possesses the (two-body) radiative decay channel: $N_1 \\to \\nu + \\gamma$, the emitted photon would carry the energy $E_\\gamma = \\mdm/2$. A large flux of such photons is expected from the large concentrations of the DM sterile neutrinos, like galaxies or galaxy clusters. Recently an extensive search of the DM decay line in the region of masses $M_\\dm \\lesssim 20\\kev$ was conducted, using the data of \\emph{Chandra}~\\citep{Riemer:06,Boyarsky:06e,Abazajian:06b} and \\emph{XMM-Newton}~\\citep{Boyarsky:05,Boyarsky:06b,Boyarsky:06c,Watson:06,Boyarsky:06d}. The region of soft X-ray (down to energies $0.2$~keV) was explored by~\\citet{Boyarsky:06f} with the use of the wide field of view spectrometer~\\citep{McCammon:02}. The non-observation of the DM decay line in X-ray, combined with the first principles calculation of DM production in the early Universe~\\citep{Asaka:06c}, implies that the \\citet{Dodelson:93} (DW) scenario can work only if the sterile neutrino mass is below 4~keV~\\citep{Boyarsky:07a}. If one takes into account recent \\emph{lower} bound on the mass of sterile neutrino DM in the DW scenario $\\mdm \\ge 5.6\\kev$~\\citep{Viel:07}, it seems that the possibility that all the DM is produced via DW scenario is ruled out~\\citep{Boyarsky:07a}. The possibility that only fraction of the DM is produced via DW mechanism remains open~\\citep{Palazzo:07}. There are other viable mechanisms of DM production, including e.g. resonant oscillation production in the presence of lepton asymmetries~\\citep{Shi:98}. Sterile neutrino DM can be produced by the decay of light inflaton \\citep{Shaposhnikov:06} or in a similar model with the different choice of parameters~\\citep{Kusenko:06a,Petraki:07}. These mechanisms are currently not constrained and remain valid for DM particles with the masses in the keV range and above. The search for the DM decay line signal produced by sterile neutrinos with masses above $\\sim 20$~keV is complicated by the absence of the focusing optics telescopes (similar to {\\it Chandra} or {\\it XMM-Newton}) in the hard X-ray and $\\gamma$-ray domain of the spectrum. For example, the existing restrictions in the $20-100$~keV mass range \\citep{Boyarsky:05,Boyarsky:06c} are derived from the observations of diffuse X-ray background, with the help of non-imaging instruments, HEAO-I~\\citep{Gruber:99}. The current status of astrophysical observations in summarized in~\\cite{Ruchayskiy:07}. In this paper we use the spectrometer SPI on board of INTEGRAL satellite to place restrictions on parameters of decaying DM in the mass range $40\\kev - 14\\mev$. This range of masses is interesting, for example, the sterile neutrinos, produced in the early Universe in the presence of large lepton asymmetries~\\citep{Shi:98} or through the inflaton decay~\\citep{Shaposhnikov:06}. It is also relevant for the case of gravitino DM~\\citep{Pagels:82,Bond:82}. When the preparation of this paper was at its final stage,~\\citet[hereafter \\textbf{Y07}]{Yuksel:07} published their work, which used the results of~\\citet[hereafter \\textbf{T06}]{Teegarden:06} to place restrictions on the parameters of sterile neutrino DM in the range $40-700\\keV$. We discuss it in more details in Section~\\ref{sec:discussion}. \\subsubsection*{SPI spectrometer} The absence of the focusing optics significantly reduces the sensitivity of the telescopes operating in the hard X-ray/soft \\gr\\ energy band. Most of the instruments operating in this energy band use collimators and/or coded masks to distinguish signals from the sources on the sky from the instrumental background. Contrary to the focusing optics telescopes, both the source and background signals are collected from the entire detector, which significantly increases the irreducible background. The focusing optics enables to significantly reduce the background only in the studies of point sources. If the source under investigation occupies a large fraction of the sky (e.g. the entire Milky Way galaxy), the performance of the focusing and non-focusing instruments with the same detector collection area are, in fact, comparable. \\begin{figure} \\centering \\includegraphics[width=\\linewidth]{SENSITIVITY_SPI} % \\caption{Comparison of sensitivity towards the search of the narrow DM decay line for different instruments with the wide FoV. Diagonal straight lines show the improvement of sensitivity (by a factor, marked on the line) as compared with the HEAO-I A4 low energy detector (LED), taken as a reference.} \\label{fig:spi_vs_heao} \\end{figure} In the case of an extended source, emitting a narrow spectral line, an efficient way of reduction of instrumental background is via the improvement of the spectral resolution of the instrument (in the case of a broad continuum background spectrum, the number of background counts at the energy of the line is proportional to the spectral resolution $\\Delta E$). The best possible sensitivity is achieved when the spectral resolution reaches the intrinsic width of the spectral line (see Fig.\\ref{fig:spi_vs_heao} for the case of wide FoV instruments and~\\cite{Boyarsky:06f} for the case of narrow FoV instruments). In the case of the line produced by the DM decaying in the Milky Way halo, the line width is determined by the Doppler broadening by the random motion of the DM particles. The velocity dispersion of the DM motion in the halo is about the rotation velocity of the Galactic disk, $v\\sim 200$~km/s. This means that Doppler broadening of the DM decay line is about \\begin{equation} \\label{eq:7} \\frac{\\Delta E}{E}\\sim \\frac{v}{c}\\simeq 10^{-3}\\;. \\end{equation} Thus, the optimal spectral resolution of an instrument searching for the DM decay line produced by the Milky Way DM halo should be $\\Delta E\\simeq 10^{-3}E$. Such optimal spectral resolution is almost achieved with the spectrometer SPI on board of \\intgr\\ satellite, which has the maximal spectral resolving power of $E/\\Delta E\\simeq 500$ and works in the energy range 20~keV -- 8~MeV~\\citep{spi}. SPI is a ``coded mask'' type instrument with an array of 19 hexagonal shaped Ge detectors (of which only 17 are operating at the moment). \\begin{figure} \\begin{center} \\includegraphics[width=\\linewidth]{FoV} \\end{center} \\caption{The geometry of the SPI FoV.} \\label{fig:FoV} \\end{figure} The SPI telescope consists of a coded mask inscribed into a circle of the radius $R_{\\rm mask}=39$~cm, placed at the height $H=171$~cm above the detector plane and of the detector, which has the shape of a hexagon inscribed into a circle of the radius $R_{\\rm det}\\simeq 15.3$~cm (see Fig. \\ref{fig:FoV}). The portion of the sky visible from each point of the SPI detector (the so-called \\emph{fully coded field of view}, FCFOV) has therefore angular diameter \\begin{equation} \\label{eq:12} \\Theta_\\fcfov=2\\arctan\\left[\\frac{R_{\\rm mask}-R_{\\rm det}}H\\right] \\approx 16^\\circ\\;, \\end{equation} while the portion of the sky visible by at least some of the detectors (the \\emph{partially coded field of view}, PCFOV) is \\begin{equation} \\label{eq:14} \\Theta_\\pcfov=2\\arctan\\left[\\frac{R_{\\rm mask}+R_{\\rm det}}H\\right] \\approx 35^\\circ\\;. \\end{equation} The solid angle spanned by the cone with this opening angle is $\\Omega_\\pcfov=2\\pi\\Bigl(1-\\cos(\\Theta_\\pcfov/2)\\Bigr)\\simeq 0.29$ (see Fig.~\\ref{fig:FoV}). Wide field of view makes the SPI telescope suitable for the study of the very extended sources, like the Milky Way DM halo. \\begin{figure} \\begin{center} \\includegraphics[width=\\linewidth]{Aeff} \\end{center} \\caption{The effective area of the SPI detector for an on-axis source, as a function of the photon energy. The plot is produced by collective the on-axis effective areas of the 17 SPI detectors from the instrumental characteristics files.} \\label{fig:Aeffon} \\end{figure} ", "conclusions": "\\label{sec:discussion} The purpose of this work was to understand how to search for the DM decay line with the SPI spectrometer and to check that none of the strong lines, present in the SPI background, was confused with the DM decay line. Our analysis shows that all the strong lines were, indeed, of instrumental origin and provides the upper bound on the flux of ``weak'' ($3{-}4\\sigma$ above the background) lines, which leads to the corresponding restrictions (see Sec.~\\ref{sec:exclusions}). To further improve the results, one needs to work with the weak lines (or lines, coinciding in position with instrumental ones). To do this one needs more sophisticated procedures of subtraction of the instrumental background (e.g. imaging). One of the most interesting cases of the coinciding instrumental and celestial line is the positronium annihilation line at 511 keV. An excess of positron annihilation emission on top of the strong instrumental line (related to positrons annihilating inside the detector) was noticed long ago~\\citep[for an incomplete set of references see e.g.][]{Prantzos:93,Milne:99,Cheng:97,Purcell:97,Knodlseder:05,Weidenspointner:06,Weidenspointner:07}. There exist many attempts of explanation of this excess. In particular, it was attributed to the annihilating or decaying DM~\\citep[see e.g.][]{Boehm:03,Hooper:03,Boehm:06,Frere:06,Picciotto:04,Rasera:05}. The sterile neutrino DM with the mass $m_s > 1\\mev$ possesses decay channel $N_s\\to e^+e^-\\nu$, with positrons annihilating either in flight or at rest, by forming the positronium atom~\\citep[see e.g.][]{Beacom:06,Sizun:06}. Thus, it is possible that the decay of sterile neutrino DM contributes to such a line. The detailed analysis of this case will be reported separately. It should be also mentioned, that the region of masses between $20\\kev \\lesssim m_\\dm \\lesssim 40\\kev$ remains inaccessible for the existing X-ray missions. The strongest restrictions in this region were produced, using the data of HEAO-1 mission \\citep{Boyarsky:06c}. When the work on this paper was at its final stage, the work of Y07 was published. Y07 obtained the restrictions on parameters of sterile neutrino in the range 40 keV -- 700 keV. To facilitate the comparison, we plot the restrictions of Y07 on Fig.~\\ref{fig:on_limit}, (divided by the factor of 2 to translate them into the restrictions for the Majorana, rather than Dirac sterile neutrino DM, see footnote~\\ref{fn:1}, p~\\pageref{fn:1}). As the data, used in our work, has about 5 times longer exposure than the \\intgr\\ first years data, on which the results of Y07 are based, we could have expected results stronger by a factor $\\approx 2$ in our case. However, the Fig.~\\ref{fig:on_limit} shows the opposite. The reason for this is as follows. For the SPI, the sensitivity towards the line search from a particular source depends on the shape of the source. In particular, the results of TW06, on which the work of Y07 was based, were obtained under the assumption of a particular diffuse source ($10^\\circ$ Gaussian). As any realistic DM profile is much flatter than the $10^\\circ$ Gaussian, the results of TW06 cannot be applied directly for the case of the DM line search. They should be rescaled to account for the diffuse nature of the DM source~(c.f. Section~\\ref{sec:exclusions}). Apart from this, the estimated DM signal from the inner part of the Galaxy is about 2 times stronger in Y07 than in our work. As the DM signal in the direction of the GC is the most uncertain, we have adopted the conservative flat profile everywhere inside the solar radius, to minimize this uncertainty. \\subsection*" }, "0710/0710.1500_arXiv.txt": { "abstract": "Observations of the ELAIS-N1 field taken at 610~MHz with the Giant Metrewave Radio Telescope are presented. Nineteen pointings were observed, covering a total area of $\\sim9$~deg$^{2}$ with a resolution of 6~$\\times$~5~arcsec$^{2}$, PA $+45\\degr$. Four of the pointings were deep observations with an rms of $\\sim40$~$\\mu$Jy before primary beam correction, with the remaining fifteen pointings having an rms of $\\sim70~\\mu$Jy. The techniques used for data reduction and production of a mosaicked image of the region are described, and the final mosaic is presented, along with a catalogue of 2500 sources detected above 6$\\sigma$. This work complements the large amount of optical and infrared data already available on the region. We calculate 610-MHz source counts down to 270~$\\mu$Jy, and find further evidence for the turnover in differential number counts below 1~mJy, previously seen at both 610~MHz and 1.4~GHz. ", "introduction": "The {\\it Spitzer} Wide-area Infrared Extragalactic \\citep[SWIRE;][]{Lonsdale03} survey has the largest sky coverage of the legacy surveys being performed by the {\\it Spitzer Space Telescope} \\citep{Werner04}. A total area of $\\sim$49~deg$^{2}$ of sky has been observed with the Infrared Array Camera \\citep[IRAC;][]{Fazio04} and Multiband Imaging Photometer for {\\it Spitzer} \\citep[MIPS;][]{Rieke04} instruments at 3.6, 4.5, 5.8, 8, 24, 70 and 160~$\\mu$m. The survey is broken down into six fields, three in the northern sky -- ELAIS-N1, ELAIS-N2 and the Lockman Hole -- and three in the south -- ELAIS-S1, {\\it Chandra} Deep Field South and the {\\it XMM}-Large Scale Structure (XMM-LSS) field. All six regions were selected to be away from the Galactic disk, in order to minimize background cirrus emission. There is a large amount of multi-wavelength information available on all six SWIRE fields. The three ELAIS fields were observed as part of the European Large-Area {\\it ISO} Survey \\citep{Oliver00}, which also included another northern (-N3) and southern (-S2) field. The {\\it Infrared Space Observatory} ({\\it ISO}) observed these regions at 6.7, 15, 90 and 175~$\\mu$m, and a large number of followup observations were carried out in the optical, infrared and radio bands. A band-merged catalogue, containing the {\\it ISO} data, along with $U$, $g'$, $r'$, $i'$, $Z$, $J$, $H$ and $K$-band detections, and radio observations at 1.4~GHz has been produced -- for more details, see \\citet{RowanRobinson04}, and references therein. Observations of the ELAIS-N1 region were taken with {\\it Spitzer} in 2004 January, covering $\\sim9$~deg$^{2}$ with the IRAC and MIPS instruments. The source catalogues have been produced, and are available online \\citep{Surace04}, containing over 280,000 sources. The UK Infrared Deep Sky Survey \\citep[UKIDSS;][]{Lawrence07} intends to cover the ELAIS-N1 region in its Deep Extragalactic Survey plan, observing the full field in the $J$, $H$ and $K$-bands to a depth of $K$ = 21~mag. This will be a great improvement over the currently available surveys, which have a sensitivity limit of $\\sim$18~mag in the $K$ band. Data Release 2 \\citep{Warren07} of UKIDSS contains early shallow data on the ELAIS-N1 region. Further surveys have been carried out in the $R$-band \\citep{Fadda04}, in H$\\alpha$ \\citep{Pascual01}, and with the {\\it Chandra} X-ray telescope \\citep{Manners03,Franceshini05}. There have been several redshift surveys of the region \\citep{Trichas06,Berta07}, and the ELAIS-N1 region was also partially covered by the Sloan Digital Sky Survey \\citep[SDSS;][]{AdelmanMcCarthy07}. While there have been a great number of observations of the ELAIS-N1 region at optical and infrared wavelengths, there is comparatively little radio information available. The existing VLA 1.4~GHz survey of the three northern ELAIS fields \\citep{Ciliegi99}, which has been included into the band-merged catalogue of \\citet{RowanRobinson04}, reaches a 5$\\sigma$ limit of 0.135~mJy over 0.12~deg$^{2}$ but only a 1.15~mJy limit over its full coverage area of 4.22~deg$^{2}$. The NVSS \\citep{Condon98} and FIRST \\citep{Becker95} surveys both cover the ELAIS-N1 region, but only to relatively shallow $5\\sigma$ limits of 2.25 and 0.75~mJy respectively. A recent study of polarised compact sources \\citep{Taylor07} at 1420~MHz is underway, using the Dominion Radio Astrophysical Observatory Synthesis Telescope (DRAO ST) centered on $16^{\\rm h}11^{\\rm m}00^{\\rm s}$, $+55\\degr00'00''$ and covering 7.4~deg$^{2}$. The first 30~per~cent of observations have been completed, with maps in Stokes I, Q and U being produced with a maximum sensitivity of 78~$\\mu$Jy~beam$^{-1}$, although with a resolution of $\\sim$1~arcmin$^{2}$. In order to extend the information on this region, a much larger deep radio survey is required. In this paper, we present observations of the ELAIS-N1 survey field taken at 610~MHz with the Giant Metrewave Radio Telescope \\citep[GMRT;][]{Ananthakrishnan05}, covering $\\sim9$~deg$^{2}$ of sky with a resolution of $6\\times5$~arcsec$^{2}$, PA~$+45\\degr$, centred on $16^{\\rm h}11^{\\rm m}00^{\\rm s}$, $+55\\degr00'00''$ (J2000 coordinates, which are used throughout this paper). This survey, in combination with the deep {\\it Spitzer} data, will be used to study the infrared/radio correlation for star-forming systems \\citep[e.g.][]{Appleton04}, and the link between the triggering of star formation and AGN activity, as well as the properties of the faint radio population at 610~MHz. In Section~\\ref{sec:observations} we describe the observations and data reduction techniques used in the creation of the survey. Section~\\ref{sec:results} presents the mosaic and a source catalogue containing 2500 sources above 6$\\sigma$, along with a sample of extended sources. In Section~\\ref{sec:sourcecounts} we construct the 610~MHz differential source counts, and compare them to previous works. ", "conclusions": "\\label{sec:results} \\begin{figure} \\includegraphics[width=8cm]{EN1NOISE.PS} \\caption{The rms noise of the final mosaic, calculated using Source Extractor. The grey-scale ranges between 40 and 350~$\\mu$Jy, and the contours are at 60 and 120~$\\mu$Jy respectively.} \\label{fig:EN1noise} \\end{figure} Source Extractor \\citep[SExtractor;][]{Bertin96} was used to calculate the rms noise $\\sigma$ across the mosaic. A grid of $16\\times16$ pixels was used in order to track changes in the local noise level, which varies significantly near the brightest sources. Fig.~\\ref{fig:EN1noise} illustrates the local noise, with the grey-scale varying between 40~$\\mu$Jy (the noise level in the centre of the deep pointings) and 350~$\\mu$Jy (the noise level for the shallow pointings, at the distance where the GMRT primary beam gain was 20~per~cent of its central value). The 60 and 120~$\\mu$m contours are plotted, which cover the majority of the deep and shallow survey regions respectively. A sample region of the ELAIS-N1 survey is shown in Fig.~\\ref{fig:EN1sample} to illustrate the quality of the image. Most of the sources in our survey are unresolved, although there are several with extended structures. We present a sample of these in Fig.~\\ref{fig:extended}. \\begin{figure*} \\includegraphics[width=16cm]{EN1area.eps} \\caption{A $70\\times70$~arcmin$^{2}$ region of the 610~MHz image, to illustrate the quality of the survey. The region is located within the deeper area of the survey, and most sources are unresolved. The grey-scale ranges between $-0.2$ and 1~mJy~beam$^{-1}$, and the noise is relatively uniform and between 40 and 60~$\\mu$Jy, apart from small regions near bright sources where the noise increases.} \\label{fig:EN1sample} \\end{figure*} \\begin{figure*} \\centerline{\\subfigure[GMRTEN1~J161137.8$+$555955] {\\includegraphics[width=5cm]{IMG12.PS}} \\subfigure[GMRTEN1~J160640.9$+$560136] {\\includegraphics[width=5cm]{IMG13.PS}} \\subfigure[GMRTEN1~J161530.7$+$545231] {\\includegraphics[width=5cm]{IMG02.PS}}} \\centerline{\\subfigure[GMRTEN1~J160929.9$+$552444 and J160931.09$+$552503] {\\includegraphics[width=5cm]{IMG04.PS}} \\subfigure[GMRTEN1~J161858.8$+$545227, J161900.3$+$545305 and J161903.3$+$545240] {\\includegraphics[width=5cm]{IMG05.PS}} \\subfigure[GMRTEN1~J161148.4$+$550049, J161151.5$+$550053 and J161154.4$+$550057] {\\includegraphics[width=5cm]{IMG07.PS}}} \\centerline{\\subfigure[GMRTEN1~J160808.7$+$544723, J160809.5$+$544659, J160810.3$+$544633 and J160810.7$+$544645] {\\includegraphics[width=5cm]{IMG08.PS}} \\subfigure[GMRTEN1~J161027.4$+$541246] {\\includegraphics[width=5cm]{IMG09.PS}} \\subfigure[GMRTEN1~J161757.1$+$545110] {\\includegraphics[width=5cm]{IMG03.PS}}} \\centerline{\\subfigure[GMRTEN1~J161140.5$+$554703 and J161142.8$+$554727] {\\includegraphics[width=5cm]{IMG14.PS}} \\subfigure[GMRTEN1~J160634.8$+$543456] {\\includegraphics[width=5cm]{IMG15.PS}} \\subfigure[GMRTEN1~J160334.6$+$542900 and J160333.1$+$542914] {\\includegraphics[width=5cm]{IMG16.PS}}} \\caption{A selection of extended objects in the ELAIS-N1 610~MHz GMRT survey -- contours are plotted at $\\pm200~\\mu$Jy $\\times$ 1, $\\sqrt{2}$, 2, $2\\sqrt{2}$, 4$\\ldots$, with the exception of object (e), where the contours start at $\\pm$500~$\\mu$Jy. Negative contours are represented by dashed lines. The resolution of the beam is shown in the bottom left of each image, and the designations of each source component are given below.} \\label{fig:extended} \\end{figure*} Our GMRT data suffer from dynamic range problems near the brightest sources, and the final mosaic has increased noise and residual sidelobes in these regions. We had fewer problems with our survey of the xFLS region, due to the longer time spent on each pointing and the correspondingly better {\\it uv} coverage. There were also fewer bright sources in the xFLS field, so a much smaller region was affected by residual sidelobes. Fig.~\\ref{fig:BrightSource} shows an area around one of the bright sources, to illustrate the problems caused by the residual sidelobes. While the local noise calculated by SExtractor increases due to these residuals, some of them still have an apparent signal-to-noise level that is greater than 6. We therefore opted for a two-stage selection criteria for our final catalogue. \\begin{figure} \\includegraphics[width=8cm]{BrightSource.ps} \\caption{Source GMRTEN1~J161212.4$+$552308, and the errors surrounding it. The grey-scale ranges between $-0.2$ and 1~mJy~beam$^{-1}$, and the source has a peak brightness of 389~mJy. The region affected by an overdensity of sources is shown by the black circle, of radius 10~arcmin -- see text for more details.} \\label{fig:BrightSource} \\end{figure} \\subsection{Source fitting} An initial catalogue of 4767 sources was created using SExtractor. The mosaic was cut off at the point where the primary beam correction dropped to 20~per~cent of its central value (a radius of 0\\fdg53 from the outer pointings), however only sources inside the 30~per~cent region (0\\fdg47) were included in the catalogue to avoid the mosaic edges from affecting the estimation of local noise. The requirements for a source to be included were that it had at least 5 connected pixels with brightness greater than 3$\\sigma$, and a peak brightness greater than 6$\\sigma$. The image pixel size meant that the beam was reasonably oversampled, so the source peak was taken to be the value of the brightest pixel within a source. The integrated flux density was calculated using the FLUX\\_AUTO option within SExtractor. This creates an elliptical aperture around each object \\citep[as described in][]{Kron80}, and integrates the flux contained within the ellipse. Comparisons between the flux density obtained through this method and through the method developed in \\citet{Garn07} -- pixels above a given threshold were summed, then empirically corrected for the elliptical beam shape -- give good agreement between the two techniques. Sources with Kron flux density above 1~mJy showed no statistical difference between the two flux density measurements, with an uncertainty of 3~per~cent. Sources below 1~mJy had a Kron flux density that was systematically larger than the \\citet{Garn07} method by 6~per~cent, with an uncertainty of 4~per~cent. We chose to use the Kron method in order to avoid the empirical correction factor. In order to estimate the area affected by artefacts near the bright sources, we calculated the number of (potentially spurious) sources in a series of concentric rings centred on a bright source, and converted this to an effective source density as a function of distance. While the noise does vary across the map, the variation is smooth and relatively slow away from the bright sources. Using the more distant rings we calculated a mean source sky density $\\mu$, and an estimate of the error in this value, $\\sigma_{\\mu}$, which will be unaffected by the presence of the spurious sources, and will vary slightly between each bright central source due to the changing properties of the image. Fig.~\\ref{fig:spurioussources} shows the source density (in arbitrary units) away from the source in Fig.~\\ref{fig:BrightSource}. The clear over-density of sources can be seen within 10~arcmin of the source. If there is a significant peak in the density (greater than $\\mu + 6\\sigma_{\\mu}$), then we define the size of the affected region by finding the first radius at which the source density drops below $\\mu + 3\\sigma_{\\mu}$ -- for this source, the affected radius is 10~arcmin. The affected region has been added to Fig.~\\ref{fig:BrightSource}. Within this region, only sources with a peak brightness greater than 12$\\sigma$ are included in our catalogue -- this value was determined empirically. The source density plot for a 10~mJy source is shown in Fig.~\\ref{fig:FaintSource}, on the same scale as Fig.~\\ref{fig:spurioussources} -- while there may still be a slight overdensity near the centre, the increased noise level in this region, along with the $6\\sigma$ cut-off reduces the risk of selecting a spurious source near sources with weak over-densities. \\begin{figure} \\includegraphics[width=8cm]{Bright.eps} \\caption{Density of sources, in concentric rings around the bright source shown in Fig.~\\ref{fig:BrightSource}. The overdensity of sources near to the bright (389~mJy) central object can be clearly seen. The mean source density, far from the central source, is given by the large dashed line while the cutoff density defining the affected region is given by short dashes -- see text for more details.} \\label{fig:spurioussources} \\end{figure} \\begin{figure} \\includegraphics[width=8cm]{Faint.eps} \\caption{Density of sources, in concentric rings around a fainter (10~mJy) object in the ELAIS-N1 survey field. There is still a slight overdensity near the central source, but the signal-to-noise cutoff of $6\\sigma$ ensures that spurious sources are not included in the catalogue.} \\label{fig:FaintSource} \\end{figure} This analysis was repeated for all sources with a peak greater than 10~mJy in order to filter spurious sources. The final catalogue contains 2500 sources -- we have erred on the side of caution in order to produce a catalogue with little contamination from spurious sources. The size of the affected region is correlated with the peak brightness of a source (with Pearson product-moment correlation coefficient of 0.53), and the number of spurious sources is also correlated with the peak brightness, with correlation coefficient 0.73. The precise size of the affected region depends on the {\\it uv} coverage for the relevant pointing, the time spent on observations and the local noise levels. Table $\\ref{tab:catalogue}$ presents a sample of 60 entries in the catalogue, which is sorted by right ascension. The full table is available via {\\tt http://www.mrao.cam.ac.uk/surveys/}. Column~1 gives the IAU designation of the source, in the form GMRTEN1~Jhhmmss.s$+$ddmmss, where J represents J2000.0 coordinates, hhmmss.s represents right ascension in hours, minutes and truncated tenths of seconds, and ddmmss represents the declination in degrees, arcminutes and truncated arcseconds. Columns~2 and 3 give the right ascension and declination of the source, calculated by first moments of the relevant pixel brightnesses to give a centroid position. Column~4 gives the brightness of the peak pixel in each source, in mJy~beam$^{-1}$, and column~5 gives the local rms noise in $\\mu$Jy~beam$^{-1}$. Columns~6 and 7 give the integrated flux density and error, calculated from the local noise level and source size. Columns 8 and 9 give the $X$, $Y$ pixel coordinates of the source centroid from the mosaic image. Column 10 is the Source Extractor deblended object flag -- 1 where a nearby bright source may be affecting the calculated flux, 2 where a source has been deblended into two or more components from a single initial island of flux, and 3 when both of the above criteria apply. There are 232 sources present in our catalogue with non-zero deblend flags; it is necessary to examine the images to distinguish between the case where one extended object has been represented by two or more entries, and where two astronomically distinct objects are present. \\begin{table*} \\label{tab:catalogue} \\caption{A sample of 60 entries from the 610-MHz ELAIS-N1 catalogue, sorted by right ascension. The full version of this table is available as Supplementary Material through the online version of this article, and via {\\tt http://www.mrao.cam.ac.uk/surveys/}.} \\begin{tabular}{cccccccccc} \\hline Name & RA & Dec.\\ & Peak & Local~Noise & Int.\\ Flux Density & Error & $X$ & $Y$ & Flags\\\\ & J2000.0 & J2000.0 & mJy~beam$^{-1}$ & $\\mu$Jy~beam$^{-1}$ & mJy & mJy & & & \\\\ (1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) \\\\ \\hline GMRTEN1~J160319.2$+$542543 & 16:03:19.22 & $+$54:25:43.6 & 0.496 & 75 & 0.415 & 0.067 & 6599 & 2946 & 0\\\\ GMRTEN1~J160319.2$+$553149 & 16:03:19.24 & $+$55:31:49.1 & 0.469 & 66 & 0.383 & 0.054 & 6527 & 5589 & 0\\\\ GMRTEN1~J160319.3$+$554950 & 16:03:19.38 & $+$55:49:50.2 & 0.514 & 85 & 0.265 & 0.054 & 6506 & 6309 & 0\\\\ GMRTEN1~J160320.8$+$550645 & 16:03:20.89 & $+$55:06:45.9 & 2.617 & 97 & 3.425 & 0.175 & 6545 & 4587 & 0\\\\ GMRTEN1~J160321.1$+$553654 & 16:03:21.13 & $+$55:36:54.2 & 0.418 & 69 & 0.309 & 0.076 & 6510 & 5792 & 0\\\\ GMRTEN1~J160322.6$+$554320 & 16:03:22.67 & $+$55:43:20.8 & 0.560 & 75 & 0.444 & 0.068 & 6495 & 6049 & 0\\\\ GMRTEN1~J160323.1$+$543737 & 16:03:23.13 & $+$54:37:37.3 & 1.920 & 74 & 2.256 & 0.119 & 6564 & 3421 & 0\\\\ GMRTEN1~J160324.0$+$550618 & 16:03:24.05 & $+$55:06:18.2 & 0.635 & 84 & 0.495 & 0.079 & 6527 & 4568 & 0\\\\ GMRTEN1~J160326.0$+$540817 & 16:03:26.00 & $+$54:08:17.1 & 0.595 & 94 & 0.486 & 0.081 & 6579 & 2248 & 0\\\\ GMRTEN1~J160327.7$+$552647 & 16:03:27.78 & $+$55:26:47.3 & 3.240 & 84 & 3.867 & 0.149 & 6484 & 5387 & 0\\\\ GMRTEN1~J160327.9$+$543326 & 16:03:27.99 & $+$54:33:26.0 & 0.452 & 75 & 0.338 & 0.064 & 6540 & 3253 & 0\\\\ GMRTEN1~J160329.5$+$540705 & 16:03:29.55 & $+$54:07:05.7 & 0.869 & 131 & 0.681 & 0.117 & 6559 & 2200 & 0\\\\ GMRTEN1~J160330.8$+$542454 & 16:03:30.83 & $+$54:24:54.3 & 0.483 & 78 & 0.837 & 0.097 & 6533 & 2912 & 0\\\\ GMRTEN1~J160331.1$+$554327 & 16:03:31.14 & $+$55:43:27.2 & 0.473 & 76 & 0.555 & 0.071 & 6447 & 6052 & 0\\\\ GMRTEN1~J160331.3$+$545000 & 16:03:31.39 & $+$54:50:00.7 & 0.408 & 67 & 0.322 & 0.055 & 6503 & 3916 & 0\\\\ GMRTEN1~J160332.3$+$553000 & 16:03:32.38 & $+$55:30:00.2 & 0.508 & 69 & 0.536 & 0.076 & 6454 & 5514 & 0\\\\ GMRTEN1~J160332.6$+$554622 & 16:03:32.66 & $+$55:46:22.6 & 3.212 & 76 & 4.152 & 0.153 & 6435 & 6169 & 0\\\\ GMRTEN1~J160332.9$+$541746 & 16:03:32.90 & $+$54:17:46.8 & 0.474 & 73 & 0.381 & 0.060 & 6528 & 2627 & 0\\\\ GMRTEN1~J160333.1$+$542914 & 16:03:33.11 & $+$54:29:14.2 & 2.377 & 73 & 9.697 & 0.226 & 6515 & 3085 & 3\\\\ GMRTEN1~J160333.6$+$552623 & 16:03:33.69 & $+$55:26:23.7 & 0.773 & 78 & 0.878 & 0.101 & 6451 & 5370 & 0\\\\ GMRTEN1~J160333.7$+$540540 & 16:03:33.70 & $+$54:05:40.6 & 7.541 & 294 & 9.799 & 0.552 & 6536 & 2142 & 0\\\\ GMRTEN1~J160334.6$+$542900 & 16:03:34.68 & $+$54:29:00.2 & 2.326 & 80 & 4.235 & 0.167 & 6506 & 3075 & 3\\\\ GMRTEN1~J160335.1$+$551534 & 16:03:35.19 & $+$55:15:34.1 & 0.462 & 74 & 0.333 & 0.066 & 6454 & 4937 & 0\\\\ GMRTEN1~J160335.2$+$555419 & 16:03:35.24 & $+$55:54:19.5 & 0.762 & 105 & 0.984 & 0.108 & 6412 & 6486 & 0\\\\ GMRTEN1~J160335.2$+$540515 & 16:03:35.25 & $+$54:05:15.3 & 38.494 & 383 & 50.351 & 0.961 & 6528 & 2125 & 0\\\\ GMRTEN1~J160336.5$+$555147 & 16:03:36.55 & $+$55:51:47.5 & 0.500 & 79 & 0.662 & 0.082 & 6408 & 6385 & 0\\\\ GMRTEN1~J160336.9$+$544120 & 16:03:36.90 & $+$54:41:20.7 & 0.532 & 68 & 0.853 & 0.092 & 6480 & 3568 & 0\\\\ GMRTEN1~J160337.6$+$545944 & 16:03:37.64 & $+$54:59:44.2 & 0.547 & 80 & 0.530 & 0.077 & 6457 & 4303 & 0\\\\ GMRTEN1~J160338.7$+$554348 & 16:03:38.77 & $+$55:43:48.1 & 0.771 & 80 & 1.743 & 0.135 & 6404 & 6065 & 0\\\\ GMRTEN1~J160339.0$+$545943 & 16:03:39.05 & $+$54:59:43.8 & 0.500 & 80 & 0.415 & 0.065 & 6449 & 4303 & 0\\\\ GMRTEN1~J160339.3$+$551352 & 16:03:39.37 & $+$55:13:52.0 & 0.516 & 78 & 0.414 & 0.063 & 6432 & 4868 & 0\\\\ GMRTEN1~J160339.6$+$542953 & 16:03:39.68 & $+$54:29:53.2 & 0.701 & 66 & 0.647 & 0.078 & 6476 & 3110 & 0\\\\ GMRTEN1~J160339.7$+$550600 & 16:03:39.75 & $+$55:06:00.4 & 0.584 & 80 & 0.418 & 0.069 & 6438 & 4554 & 0\\\\ GMRTEN1~J160340.1$+$545127 & 16:03:40.10 & $+$54:51:27.4 & 0.711 & 79 & 0.639 & 0.079 & 6451 & 3972 & 0\\\\ GMRTEN1~J160340.1$+$550544 & 16:03:40.15 & $+$55:05:44.9 & 0.495 & 81 & 0.392 & 0.069 & 6436 & 4543 & 0\\\\ GMRTEN1~J160340.8$+$554325 & 16:03:40.83 & $+$55:43:25.8 & 1.447 & 81 & 4.633 & 0.210 & 6392 & 6050 & 0\\\\ GMRTEN1~J160341.2$+$552611 & 16:03:41.27 & $+$55:26:11.8 & 0.475 & 68 & 0.460 & 0.065 & 6408 & 5361 & 0\\\\ GMRTEN1~J160341.5$+$552205 & 16:03:41.59 & $+$55:22:05.4 & 0.638 & 69 & 0.614 & 0.080 & 6411 & 5197 & 0\\\\ GMRTEN1~J160341.6$+$553203 & 16:03:41.66 & $+$55:32:03.4 & 0.462 & 69 & 0.365 & 0.062 & 6400 & 5595 & 0\\\\ GMRTEN1~J160343.1$+$540324 & 16:03:43.13 & $+$54:03:24.6 & 0.908 & 126 & 0.556 & 0.092 & 6483 & 2050 & 0\\\\ GMRTEN1~J160344.8$+$540320 & 16:03:44.84 & $+$54:03:20.4 & 1.168 & 118 & 0.856 & 0.110 & 6473 & 2047 & 0\\\\ GMRTEN1~J160345.8$+$553021 & 16:03:45.84 & $+$55:30:21.4 & 0.381 & 60 & 0.370 & 0.065 & 6378 & 5527 & 0\\\\ GMRTEN1~J160345.8$+$554238 & 16:03:45.88 & $+$55:42:38.3 & 0.653 & 76 & 2.566 & 0.156 & 6365 & 6018 & 0\\\\ GMRTEN1~J160346.5$+$552855 & 16:03:46.56 & $+$55:28:55.8 & 0.577 & 73 & 0.510 & 0.076 & 6375 & 5469 & 0\\\\ GMRTEN1~J160346.6$+$550826 & 16:03:46.62 & $+$55:08:26.0 & 0.525 & 70 & 0.416 & 0.065 & 6396 & 4650 & 0\\\\ GMRTEN1~J160348.3$+$542626 & 16:03:48.37 & $+$54:26:26.4 & 2.328 & 80 & 2.660 & 0.133 & 6429 & 2970 & 0\\\\ GMRTEN1~J160348.6$+$550124 & 16:03:48.61 & $+$55:01:24.3 & 0.477 & 78 & 0.257 & 0.050 & 6392 & 4368 & 0\\\\ GMRTEN1~J160349.2$+$554243 & 16:03:49.21 & $+$55:42:43.9 & 3.349 & 85 & 4.224 & 0.163 & 6346 & 6021 & 0\\\\ GMRTEN1~J160350.0$+$545634 & 16:03:50.06 & $+$54:56:34.6 & 0.435 & 67 & 0.371 & 0.058 & 6389 & 4175 & 0\\\\ GMRTEN1~J160350.4$+$550717 & 16:03:50.43 & $+$55:07:17.2 & 0.597 & 78 & 0.626 & 0.083 & 6376 & 4603 & 0\\\\ GMRTEN1~J160350.5$+$541302 & 16:03:50.51 & $+$54:13:02.7 & 0.770 & 80 & 0.535 & 0.068 & 6430 & 2435 & 0\\\\ GMRTEN1~J160351.1$+$555644 & 16:03:51.13 & $+$55:56:44.4 & 3.996 & 120 & 7.391 & 0.287 & 6321 & 6581 & 0\\\\ GMRTEN1~J160351.2$+$552346 & 16:03:51.24 & $+$55:23:46.1 & 0.519 & 80 & 0.640 & 0.072 & 6354 & 5262 & 0\\\\ GMRTEN1~J160351.3$+$540609 & 16:03:51.38 & $+$54:06:09.9 & 0.951 & 110 & 0.843 & 0.110 & 6432 & 2159 & 0\\\\ GMRTEN1~J160352.6$+$550012 & 16:03:52.62 & $+$55:00:12.8 & 0.510 & 71 & 0.592 & 0.076 & 6370 & 4320 & 0\\\\ GMRTEN1~J160352.8$+$542943 & 16:03:52.80 & $+$54:29:43.5 & 2.258 & 82 & 2.422 & 0.124 & 6400 & 3101 & 0\\\\ GMRTEN1~J160353.2$+$550309 & 16:03:53.29 & $+$55:03:09.9 & 0.570 & 74 & 0.500 & 0.072 & 6363 & 4438 & 0\\\\ GMRTEN1~J160355.5$+$553844 & 16:03:55.52 & $+$55:38:44.9 & 0.463 & 73 & 0.242 & 0.050 & 6314 & 5861 & 0\\\\ GMRTEN1~J160356.2$+$550126 & 16:03:56.20 & $+$55:01:26.7 & 0.472 & 77 & 0.380 & 0.063 & 6348 & 4369 & 0\\\\ GMRTEN1~J160356.2$+$550415 & 16:03:56.25 & $+$55:04:15.3 & 0.506 & 83 & 0.451 & 0.068 & 6345 & 4481 & 0\\\\ \\hline \\end{tabular} \\end{table*} \\subsection{Comparison with other surveys} In order to test the positional accuracy of our catalogue, we paired it with the 393 objects in the VLA 1.4~GHz source catalogue of \\citet{Ciliegi99}, using a pairing radius of 6~arcsec. The VLA survey covers only $\\sim25$~per~cent of our 610~MHz observations. Fig.~\\ref{fig:deltaRADEC} shows the position offset of the 263 matched sources compared with their VLA counterparts -- the distribution of offsets is approximately Gaussian, with mean offset in Right Ascension of $-0.9$~arcsec, and $-0.1$~arcsec in Declination. The standard deviations of the distribution are 0.9 and 0.7~arcsec respectively. \\begin{figure} \\includegraphics[width=8cm]{deltaRADEC.ps} \\caption{Source positions in the GMRT catalogue relative to the positions found in the VLA catalogue of \\citet{Ciliegi99}, for unique matches within 6~arcsec. Offsets are distributed in a Gaussian fashion in RA and DEC, and the ellipse corresponding to 1$\\sigma$ for the distribution is shown.} \\label{fig:deltaRADEC} \\end{figure} We repeated this analysis with data from the FIRST survey \\citep{Becker95}. The whole of our 610~MHz survey region is covered by FIRST, and even with the reduced sensitivity of FIRST ($\\sim$150~$\\mu$Jy noise), 504 pairs are found within 6~arcsec. We again find Gaussian-distributed position errors, with RA offset of $-0.1$~arcsec, standard deviation 0.4~arcsec and DEC offset of $-0.1$~arcsec, standard deviation 0.6~arcsec. We have not corrected the positions given in our catalogue, since it agrees closely with FIRST. The spectral index distribution of the matched sources from both surveys is shown in Fig.~\\ref{fig:alpha}, using the integrated flux density measurements. The distribution peaks around $\\alpha=0.7$, where $\\alpha$ is defined such that the flux density $S$ scales with frequency $\\nu$ as $S = S_{0}\\nu^{-\\alpha}$. \\begin{figure} \\includegraphics[width=8cm]{Alpha.eps} \\caption{Radio spectral index $\\alpha$ between 610~MHz and 1.4~GHz, for sources in the VLA ELAIS-N1 catalogue of Ciliegi et al.\\ (solid lines) and in the FIRST catalogue (dashed lines).} \\label{fig:alpha} \\end{figure} Fig.~\\ref{fig:AlphaFlux} shows the spectral index distribution for all sources with matches in the GMRT and FIRST catalogues (black diagonal crosses), and sources in the GMRT and \\citet{Ciliegi99} catalogues (red upright crosses). There are significant biases in Fig.~\\ref{fig:AlphaFlux}, due to the varying sensitivity levels of the three surveys. In the region covered by \\citet{Ciliegi99}, the 610~MHz completeness level is 360~$\\mu$Jy, shown by the black dotted line. The limiting spectral indices for sources at the sensitivity levels of the 1.4~GHz surveys are shown (FIRST -- solid black line, Ciliegi -- dashed red line). In order to look for variations in the source population, we calculate the mean and median spectral indices for sources with detections in the Ciliegi catalogue, with 610~MHz flux density between 500~$\\mu$Jy and 1~mJy -- the point at which the turnover in source counts becomes visible (see Section~\\ref{sec:sourcecounts}) -- and above 1~mJy. The mean values of $\\alpha$ are $0.22\\pm0.09$ and $0.45\\pm0.04$ respectively, and the median values are $0.36\\pm0.12$ and $0.56\\pm0.04$. There are 48 and 168 sources in the two flux density bins. The bias against steep-spectrum sources at low flux densities (which is visible in Fig.~\\ref{fig:AlphaFlux}) means that, for a source with 610~MHz flux density of 500~$\\mu$Jy, the largest value of $\\alpha$ that would be detectable is $\\sim$0.8 and so this apparent flattening at fainter flux densities may simply be due to sample bias. However, \\citet{Bondi07} also find significantly flatter spectral indices for fainter radio sources, again comparing 610~MHz and 1.4~GHz data, and attribute this to the emergence of a population of low-luminosity AGNs -- see Section~\\ref{sec:sourcecounts} for more details. \\begin{figure} \\includegraphics[width=8cm]{AlphaFlux.eps} \\caption{The variation in spectral index $\\alpha$ with 610~MHz flux density. Black diagonal crosses represent sources in the GMRT and FIRST catalogues, with the solid black line showing the limiting spectral index that could be detected, given the respective sensitivity levels. Red upright crosses represent sources in the corresponding GMRT and Ciliegi et al.\\ catalogues, with the dashed red line showing the limit on $\\alpha$. The 610~MHz flux density limit is shown by the dotted black line.} \\label{fig:AlphaFlux} \\end{figure}" }, "0710/0710.0506_arXiv.txt": { "abstract": "Using a combination of deep MID-IR observations obtained by IRAC, MIPS and IRS on board Spitzer we investigate the MID-IR properties of Lyman Break Galaxies (LBGs) at z$\\sim$3, establish a better understanding of their nature and attempt a complete characterisation of the population. With deep mid-infrared and optical observations of $\\sim$1000 LBGs covered by IRAC/MIPS and from the ground respectively, we extend the spectral energy distributions (SEDs) of the LBGs to mid-infrared. Spitzer data reveal for the first time that the mid-infrared properties of the population are inhomogeneous ranging from those with marginal IRAC detections to those with bright rest-frame near-infrared colors and those detected at 24$\\mu$m MIPS band revealing the newly discovered population of the Infrared Luminous Lyman Break Galaxies (ILLBGs). To investigate this diversity, we examine the photometric properties of the population and we use stellar population synthesis models to probe the stellar content of these galaxies. We find that a fraction of LBGs have very red colors and large estimated stellar masses $M_{\\ast}$$>$5$\\times$$10^{10}$$M_{\\odot}$. We discuss the link between these LBGs and submm-luminous galaxies and we report the detection of rest frame 6.2 and 7.7 $\\mu$m emission features arising from Polycyclic Aromatic Hydrocarbons (PAH) in the Spitzer/IRS spectrum of an infrared-luminous Lyman break galaxy at z=3.01. ", "introduction": "Observation and study of high-redshift galaxies is essential to constrain the history of galaxy evolution and give us a systematic and quantitative picture of galaxies in the early universe, an epoch of rigorous star and galaxy formation. Large samples of high-z galaxies that have recently become available, play a key role to that direction and have revealed a zoo of different galaxy populations at z. There are various techniques for detecting high-z galaxies involving observations in wavelengths that span from optical to far-IR. Among the various methods the Lyman break dropout technique (\\cite{Steidel93}), sensitive to the presence of the 912{\\AA} break, is designed to select z$\\sim$ 3 galaxies. LBGs constitute at the moment the largest galaxy population at z$\\sim$3 (\\cite{Steidel03}). With observations spanning from X-rays (eg. \\cite{Nan02}) to near-infrared (\\cite{Shapley03}) there has been a considerable progress into understanding the nature of population, but to fully characterize their properties (such as stellar mass, dust content, link to other z$\\sim$3 populations) observations of longer wavelengths are required. With the advent of Spitzer Space Telescope (\\cite{Werner04}) we have access to longer wavelengths. IRAC bands (3.6, 4.5, 5.8, 8.0$\\mu$m) are crucial as they trace the rest-frame near infrared luminosities for galaxies at 0.5$<$z$<$5, (where the bulk of the stellar mass of a galaxy radiates) while MIPS (24, 70, 160 $\\mu$m) and IRS (5.3--40$\\mu$m) provide an insight into the interstellar medium of the population as they are sensitive to PAH features and dust re-radiation. In this study we use IRAC and MIPS data covering $\\sim$1000 and 244 LBGs respectively, lying on the fields Q1422+2309 (Q1422), DSF2237a,b (DSF), Q2233+1341 (Q2233), SSA22a,b (SSA22), B20902+34 (B0902), QSOHS1700+6416 (Q1700), Extended Groth Strip (EGS) and Hubble Deep Field North (HDFN). Those LBGs have previously beeen identified from their optical colours by \\cite{Steidel03}. In section 2 we search for mid-infrared counterparts of the LBGs, extend their SEDs to mid-infrared and investigate their mid-infrared colours as well as their physical properties such as stellar mass and dust content. In section 3 we examine the possible link between the IRAC/MIPS bright LBGs and the SMGs while in Section 4 using data obtained by IRS, we report the detection of PAH features arising from the mid-inrared spectrum of an ILLBG at z=3.01. In Section 5 we summurize the results of this Spitzer view of LBGs. ", "conclusions": "The advent of Spitzer has dramatically improved our understanding of the LBGs. Using data obtained by IRAC/MIPS/IRS on board Spitzer we have reached the following conclusions for the population of the LBGS: \\begin{itemize} \\item IRAC colors have revealed the diversity of LBGs ranging from those with marginal detection in IRAC bands and R-[3.6]$<$1.5 colors, to those that have bright in IRAC bands and exhibit R-[3.6]$>$1.5 colors. \\item LBGs detected at 8$\\mu$m have redder R-[3.6] colors and on average are more massive, suffer more obscuration and have relatively older stellar populations when compared to the rest of the population. \\item A fraction of about $\\sim$5\\% of the LBGs do have dust as evidenced by MIPS 24$mu$m detections, and are classified as ILLBGs. Those LBGs share many properties in common with the SMGs and preliminary results show that they can be detected at submm bands. It can therefore be suggested that a link between these two populations must exist. \\item Strong PAH features arising from the mid-infrared spectrum of an ILLBG at z=3.01 indicates the existence of dust in the interstellar medium of LBGs and suggest that the emission is dominated by star formation rather than an AGN. \\end{itemize} \\begin{figure} \\centering \\includegraphics[width=10cm,height=4cm]{magdis_fig2.eps} \\caption {\\small{IRS spectrum of EGS20 J1418+5236. The dashed line is the M82 SED shifted to $z=3.01$. The spectrum has remarkably similar 6.2 and 7.7 $\\mu$m PAH emission-feature strength and shape to those of M82. We cross-correlated the IRS spectrum with that of M82 to derive a redshift of $z =3.01$$\\pm$$0.016$.}} \\end{figure}" }, "0710/0710.5609_arXiv.txt": { "abstract": "{} {We present the results of a spectroscopic analysis on three young embedded sources (HH26 IRS, HH34 IRS and HH46 IRS) belonging to different star-forming regions and displaying well developed jet structures. The aim is to investigate the source accretion and ejection properties and their connection.} {We used VLT-ISAAC near-IR medium resolution ($R\\sim9000$) spectra ($H$ and $K$ bands) to derive, in a self-consistent way, parameters like the star luminosity, the accretion luminosity and the mass accretion rate. Mass ejection rates have also been estimated from the analysis of different emission features.} {The spectra present several emission lines but no photospheric features in absorption, indicating a large veiling in both $H$ and $K$ bands. In addition to features commonly observed in jet driving sources ([\\ion{Fe}{ii}], H$_2$, \\ion{H}{i}, CO), we detect a number of emission lines due to permitted atomic transitions, such as \\ion{Na}{i} and \\ion{Ti}{i} that are only 2-5 times weaker than the Br$\\gamma$ line. Some of these features remain unidentified. Emission from \\ion{Na}{i} 2.2$\\mu$m doublet is observed along with CO(2-0) band-head emission, indicating a common origin in an inner gaseous disc heated by accretion. We find that accretion provides about 50\\% and 80\\% of the bolometric luminosity in HH26 IRS and HH34 IRS, as expected for accreting young objects. Mass accretion and loss rates spanning $10^{-6}$--$10^{-8}$ M$_{\\sun}$\\,yr$^{-1}$ have been measured. The derived $\\dot{M}_\\mathrm{loss}/\\dot{M}_\\mathrm{acc}$ is $\\sim$0.01 for HH26 IRS and HH34 IRS, and $>$0.1 for HH46 IRS. These numbers are in the range of values predicted by MHD jet launching models and found in the most active classical T Tauri stars.} {Comparison with other spectroscopic studies performed on Class Is seems to indicate that Class Is actually having accretion-dominated luminosities are a limited number. Although the analysed sample is small, we can tentatively define some criteria to characterise such sources: they have $K$-band veiling larger than 2 and in the majority of the cases present IR features of CO and \\ion{Na}{i} in emission, although these do not directly correlate with the accretion luminosity. Class Is with massive jets have high $L_{\\mathrm{acc}}/L_{\\mathrm{bol}}$ ratios but not all the identified accretion-dominated objects present a jet. As suggested by the SEDs of our three objects, the accretion-dominated objects could be in an evolutionary transition phase between Class 0 and I. Studies of the kind presented here but on larger samples of possible candidates should be performed in order to test and refine these criteria.} ", "introduction": "The process of mass accretion accompanying the formation of solar type stars is always associated with mass ejection in form of collimated jets, that extend from few AU up to parsecs distance from the exciting source. According to an established class of models \\citep{koenigl00,casse00}, accretion and ejection are indeed intimately related through the presence of a magnetised accretion disc: the jets carry away the excess angular momentum, so that part of the disc material can move toward the star. The efficiency of this process is measured by the ratio between the mass ejection and mass accretion rates, and depends on the jet acceleration mechanism at work. Measurements of such an efficiency have been so far obtained only for classical T Tauri stars, whose accretion properties are rather well studied through the observation of the excess continuum emission at optical and UV wavelengths Values of $\\dot{M}_\\mathrm{loss}/\\dot{M}_\\mathrm{acc}$ in the range 1-10\\% have been found by different studies \\citep[e.g.][]{hartigan95,woitas05,ferreira06}. It is nevertheless important to test accretion/ejection models in young sources at earlier stages of evolution, when accretion dominates the energetics of the system and thus the mechanism to extract angular momentum is expected to be more efficient. To this aim, an interesting sample of objects are Class I sources, i.e. the class of embedded stars characterised by a steeply rising IR spectral energy distribution (SED) between 2 and 10 $\\mu$m and usually considered younger than visible T Tauri stars (the Class II sources). However, the high extinction pertaining to these objects strongly limits the measurement of their stellar and accretion properties, needed to prove that they are indeed in a phase of higher accretion with respect to Class II sources. The general assumption so far applied has been that most of the bolometric luminosity of Class I objects is due to accretion. Recently, however, thanks to the use of high dispersion sensitive instrumentation, it has become possible to define the stellar properties of small samples of Class I stars through their weak photospheric lines detected in the optical scattered light \\citep{white04} and in the near-IR direct emission \\citep{greene02,nisini05a,doppmann05}. Such studies have shown that the characteristics of Class I objects vary in fact significantly. In particular, the accretion luminosity may span from few percent up to 80\\% of the bolometric luminosity; these findings show that not all sources defined as Class I are indeed actively accreting objects and suggest that a classification based on different criteria is indeed required. \\citet{white04}, notably, derived from their analysis of the scattered light spectra of Taurus-Auriga sources, that the average $\\dot{M}_\\mathrm{loss}/\\dot{M}_\\mathrm{acc}$ for Class I of their sample is larger than for Class II and close to unity. They interpret this result as due to an observational bias induced by the effect of disc orientation in their optical spectra; if the Class I are seen prevalently edge-on, then the extended region emitting the forbidden lines from which they estimate $\\dot{M}_\\mathrm{loss}$ is seen more directly than the obscured stellar photosphere. It is clear that such kind of biases can be minimised by performing spectroscopic observations directly in the IR, where features originating in the photosphere, in the accretion region and in the jet can be simultaneously detected by instrumentation which is sensitive enough. In this framework, we report here the results of near-IR spectroscopic observations at medium resolution of three embedded sources (HH34 IRS, HH26 IRS and HH46 IRS) displaying well developed jet structures and having a spectral index between 2 and 10\\,$\\mu$m, typical of Class I objects. We have derived accretion and ejection parameters of these sources through the analysis of the different features detected on the spectra. The goal is to study how much of their energy is due to accretion and to investigate the efficiency of the ejection mechanism. We describe the sample and the observations in Sec. 2 and report the results in Sec. 3; in Sec. 4 we present the procedures applied for the analysis of the data and infer the physical properties of the objects and their jets. These results are then discussed in Sec. 5, where a comparison with similar sources analysed in previous studies will be also made. Main conclusions of our work are summarised in Sec. 6. \\begin{table*}[!t] \\begin{center} \\caption[]{The observed targets and their main observational properties.} \\begin{tabular}{l c c | c | c c | c c | c c c c} \\hline \\hline Source & R.A.(2000)\t\t\t&DEC.(2000)\t\t&$m_\\mathrm{J}$& \\multicolumn{2}{|c|}{$m_\\mathrm{H}$}\t \t& \\multicolumn{2}{c|}{$m_\\mathrm{K}$}\t& $D$\t\t& $\\alpha^{(a)}$\t\t& $L_\\mathrm{bol}$\t\t\\\\ &\t\t\t\t\t\t&\t\t\t\t&(mag)\t\t\t\t\t&\\multicolumn{2}{|c|}{(mag)}&\\multicolumn{2}{|c|}{(mag)}&(pc)\t&\t&\t(L$_{\\sun}$)\t\\\\ &\t\t\t\t\t\t&\t\t\t\t&2MASS$^{(1)}$&2MASS$^{(1)}$&This work$^{(1)}$&2MASS$^{(b)}$&This work$^{(b)}$&\t&\t&\t\t\\\\ \\hline HH 26 IRS\t& 05 46 03.9\t& -00 14 52\t\t\t&16.77& 14.07 &14.6 & 11.88 & 12.3 \t\t& 450 \t\t& 2.01 \t\t\t& 4.6--9.2\t\\\\ HH 34 IRS\t& 05 35 29.9\t& -06 26 58\t \t\t&15.06& 13.60 &13.5 \t& 12.38 & 12.4 \t\t& 460 \t\t& 1.14 \t\t\t& 12.4--19.9\t \\\\ HH 46 IRS\t& 08 25 43.9\t& -51 00 36\t \t\t&14.20& 12.88 &14.8 \t& 12.72 & 13.4 \t\t& 450 \t\t& 1.96 \t\t\t& $<$15.0$^{(c)}$ \\\\ \\hline \\end{tabular} \\label{targets} \\end{center} \\vspace{0.2 cm} \\small{Notes. (a) The spectral index $\\alpha=d\\mathrm{Log}(\\lambda F_{\\lambda})/d\\mathrm{Log}(\\lambda)$ is calculated between 2 and 10 $\\mu$m. (b) The magnitude values in the $H$ and $J$ bands have been estimated from the calibrated spectra. (c) Total luminosity: the source is a binary (see text for details). References. (1) from the 2MASS catalogue \\citep{2mass}.} \\end{table*} ", "conclusions": "We have investigated the accretion and ejection properties of three embedded sources (HH26 IRS, HH34 IRS, HH46 IRS) showing prominent jet-like structures. To this aim we have analysed their medium resolution (R $\\sim$ 9000) near IR spectra acquired with VLT-ISAAC. The main results we obtained can be summarised as follows: \\begin{itemize} \\item The bolometric luminosity and SEDs of the three sources have been revised on the basis of Spitzer and recent sub-mm observations, and theoretical radiative transfer models available in the literature. It turns out that the sources are probably very young, in a transition phase between Class 0 and I. \\item The spectra of the three sources show important differences in their characteristics: in fact, the number and the absolute and relative intensity of the observed emission features (associated both with the accretion region and the jet environment) vary among the sources. In particular, we point out that there is no clear relationship between the presence of the jet and the spectral accretion signatures detected. Moreover, the spectra show no sign of absorption features, indicating large amount of veiling which in turn suggests the presence of warm dusty envelopes around the sources, heated by the active ongoing accretion. \\item In two of our sources (HH34 IRS and HH26 IRS) we find a Br$\\gamma$/\\ion{Na}{i} ratio much larger (by a factor of three or more) than that observed in other Class Is or T Tauri stars. Conversely, the ratio between the CO(2-0) 2.3$\\mu$m overtone emission and the \\ion{Na}{i} 2.20$\\mu$m doublet, shows the opposite trend, i.e. is smaller in the objects with jets. This may indicate the presence of massive discs around the jet sources, characterised by large amounts of warm gas in neutral state. \\item We consistently derive $A_K$, $L_*$ and $L_{\\mathrm{acc}}$ assuming the three sources on the birthline and adopting the relationship between Br$\\gamma$ luminosity and accretion luminosity derived by \\citet{muzerolle98}. In the case of HH34 IRS and HH26 IRS, accretion largely contributes to the total energy, as expected for young sources in the main accretion phase. In HH46 IRS, we estimate an $L_{\\mathrm{acc}}/L_{\\mathrm{bol}}$ ratio of the order of 0.2, but the real value largely depends on how the bolometric luminosity is shared among the two binary components. Taking into account $K$-band veiling values $r_K$ greater than the lower limits inferred from the spectra would lead in general to higher $A_K$, $L_{\\mathrm{acc}}$ and $\\dot{M}_{acc}$. \\item Mass accretion and loss rates span (including errors) in the range $10^{-8}$--$10^{-6}$ M$_{\\sun}$\\,yr$^{-1}$. The derived $\\dot{M}_{loss}/\\dot{M}_{acc}$ ratio is $\\sim$0.01-0.03 for HH26 IRS and HH34 IRS, and $>$0.1 for HH46 IRS. These numbers are in the range of values usually predicted by models and found in the more active classical T Tauri stars. \\item Comparing the results found in this work with other spectroscopic studies recently performed on Class I sources, we conclude that the number of Class Is actually having accretion-dominated luminosities (Accretion-Dominated Young Objects, ADYOs) could be limited. From the properties inferred in a small sample of objects we can tentatively define some criteria to characterise such sources: ADYOs have all $K$-band veiling larger than 2 and in the majority of the cases present (in addition to \\ion{H}{i}) IR features of CO and \\ion{Na}{i} in emission, although these latter do not directly correlate with the accretion luminosity. Class Is with massive jets have high $L_{\\mathrm{acc}}/L_{\\mathrm{bol}}$ ratios but not all the identified ADYOs present a jet. The SEDs of our small sample of three objects, suggest that accretion-dominated sources could be in an evolutionary phase in transition between Class 0 and I. Of course, studies of the kind presented here but carried out on larger samples of possible candidates should be performed in order to test and refine the tentative criteria that we have just mentioned. \\end{itemize}" }, "0710/0710.0440_arXiv.txt": { "abstract": "We examine the ability of a future X-ray observatory, with capabilities similar to those planned for the Constellation-X or X-ray Evolving Universe Spectroscopy (XEUS) missions, to constrain dark energy via measurements of the cluster X-ray gas mass fraction, $f_{\\rm gas}$. We find that $f_{\\rm gas}$ measurements for a sample of $\\sim500$ hot ($kT\\gsim5$keV), X-ray bright, dynamically relaxed clusters, to a precision of $\\sim 5$ per cent, can be used to constrain dark energy with a Dark Energy Task Force (DETF; Albrecht et al. 2006) figure of merit of $15-40$, with the possibility of boosting these values by 40 per cent or more by optimizing the redshift distribution of target clusters. Such constraints are comparable to those predicted by the DETF for other leading, planned `Stage IV' dark energy experiments. A future $f_{\\rm gas}$ experiment will be preceded by a large X-ray or SZ survey that will find hot, X-ray luminous clusters out to high redshifts. Short `snapshot' observations with the new X-ray observatory should then be able to identify a sample of $\\sim 500$ suitably relaxed systems. The redshift, temperature and X-ray luminosity range of interest has already been partially probed by existing X-ray cluster surveys which allow reasonable estimates of the fraction of clusters that will be suitably relaxed for $f_{\\rm gas}$ work to be made; these surveys also show that X-ray flux contamination from point sources is likely to be small for the majority of the targets of interest. Our analysis uses a Markov Chain Monte Carlo method which fully captures the relevant degeneracies between parameters and facilitates the incorporation of priors and systematic uncertainties in the analysis. We explore the effects of such uncertainties for scenarios ranging from optimistic to pessimistic. We conclude that the $f_{\\rm gas}$ experiment offers a competitive and complementary approach to the best other large, planned dark energy experiments. In particular, the $f_{\\rm gas}$ experiment will provide tight constraints on the mean matter and dark energy densities, with a peak sensitivity for dark energy work at redshifts midway between those of supernovae and baryon acoustic oscillation/weak lensing/cluster number counts experiments. In combination, these experiments should enable a precise measurement of the evolution of dark energy. ", "introduction": "\\label{introduction} In the early 1990s, measurements of the baryonic mass fraction in X-ray luminous galaxy clusters provided compelling evidence that we live in a low density Universe. Under the assumption that large clusters provide approximately fair samples of the matter content of the Universe, X-ray observations require that the mean matter density, $\\Omega_{\\rm m}$, is significantly less than the critical value, with a best-fit value $\\Omega_{\\rm m}\\sim 0.2-0.3$ \\citep[e.g.][]{White:91,Fabian:91, Briel:92, White:93, David:95, White:95, Evrard:97, Mohr:99, Ettori:99, Roussel:00, Grego:01, Allen:02, Allen:04, Allen:07, Ettori:03, Sanderson:03, Lin:03, LaRoque:06}. When combined with the expectation from inflation models, later confirmed by Cosmic Microwave Background (CMB) studies \\citep[][and references therein]{Bennett:03, Spergel:03}, that the Universe should be close to spatially flat, X-ray results on the cluster baryon mass fraction quickly lead to the suggestion that the mass-energy density of the Universe may be dominated by a cosmological constant \\citep[e.g.][]{White:93}. The first direct evidence for late-time cosmic acceleration, as would be produced by a sizeable cosmological constant, was provided in the late 1990s by \\cite{Riess:98} and \\cite{Perlmutter:98} based on measurements of the light curves of type Ia supernovae (SNIa). Since then, larger SNIa data sets \\citep{knop:03, Riess:04, Astier:06, Riess:07, WoodVasey:07, Davis:07} and an increasingly wide array of other, complementary experiments have confirmed and improved upon this striking measurement. The combination of CMB data from the Wilkinson Microwave Anisotropy Probe (WMAP) \\citep{Spergel:03, Spergel:06, Dunkley:08} with large scale structure (LSS) data from the Sloan Digital Sky Survey (SDSS) \\citep{Eisenstein:05, Percival:07} and/or 2dF Galaxy Redshift Survey (2dFGRS) \\citep{Cole:05} provides powerful evidence for dark energy. The cross-correlation of CMB and LSS fluctuations reveals the effects of dark energy on the Integrated Sachs-Wolfe effect \\citep{Scranton:03, Fosalba:03, Rassat:07}. Measurements of the amplitude and evolution of matter fluctuations using X-ray galaxy clusters \\citep{Borgani:01,Reiprich:02, Allen:03, Schuecker:03, Voevodkin:04, Henry:04, Mantz:07}, optically-selected clusters \\citep{Gladders:06, Rozo:07}, Lyman-$\\alpha$ forest data \\citep{Viel:04, Seljak:04}, and weak lensing \\citep{vanWaerbeke:05, Jarvis:05, Hoekstra:06, Benjamin:07}, also provide important, powerful confirmation of the new, standard cosmological paradigm: a universe in which the main mass and energy components are dark matter and dark energy, and where dark energy drives the current acceleration. The standard model for dark energy remains the cosmological constant, which is mathematically equivalent to vacuum energy. In principle, however, cosmic acceleration could be driven by either dark energy or a modification to the laws of gravity on cosmological scales \\citep[see][for an extensive review]{Copeland:06}. Building on the early X-ray work, \\cite{Allen:04, Rapetti:05}; and \\cite{Allen:07} showed that measurements of the evolution of the X-ray gas mass fraction, $f_{\\rm gas}$, in the largest, dynamically relaxed galaxy clusters provides a further powerful, complementary approach for studying dark energy. As with SNIa data, $f_{\\rm gas}(z)$ measurements probe the redshift-distance relation; whereas the peak SNIa luminosity varies as the square of the distance, $f_{\\rm gas}$ measurements vary as distance, $d^{1.5}$. \\citep[The distance dependance derives from the way in which $f_{\\rm gas}$ values are determined from the observed X-ray temperature and surface brightness data;][]{Allen:07} In combination with the tight constraint on $\\Omega_{\\rm m}$ provided by the normalization of the $f_{\\rm gas}(z)$ curve, under the assumption of fair matter samples, the $f_{\\rm gas}(z)$ data contain sufficient information to break the degeneracy between $\\Omega_{\\rm m}$ and the dark energy equation of state, $w$, in the distance equations. The additional combination of $f_{\\rm gas}$ and CMB data breaks other important degeneracies between parameters in cosmological analyses \\citep{Rapetti:05, Allen:07}. \\cite{Allen:07} show that the current constraints on dark energy from the $f_{\\rm gas}$ experiment are of comparable precision to other leading techniques, and are robust under the inclusion of conservative systematic allowances, e.g. relaxing the requirement for exact hydrostatic equilibrium and allowing for moderate redshift evolution in the cluster baryon fraction. These authors also show that intrinsic, systematic scatter remains undetected in the current $f_{\\rm gas}$ data, despite a weighted mean statistical scatter in the individual distance measurements of only $\\sim 5$ per cent; in contrast, SNIa studies \\citep{Riess:07, Jha:07, WoodVasey:07} have established the presence of systematic scatter at the $\\sim 7$ per cent in distance measurements from the best current SNIa data. The key to determining the nature of dark energy is to obtain precise measurements of its evolution with redshift, $z$, or scale factor, $a=1/(1+z)$. The Dark Energy Task Force report \\cite[][hereafter DETF]{Albrecht:06} presented estimates of the constraints on dark energy parameters that should be achievable with a number of future proposed or planned dark energy experiments. In particular, the report forecasted the ability of these experiments, in combination with CMB data from the Planck satellite, to constrain a dark energy model of the form $w(a)=w_{\\rm 0}+w_{\\rm a}(1-a)$, and defined a figure of merit (hereafter FoM) to allow for easy comparison of the constraints. In this paper, we use the same dark energy parameterization and FoM to quantify the constraining power of future $f_{\\rm gas}$ experiments, to be carried out with e.g. the Constellation-X or X-ray Evolving Universe Spectroscopy (XEUS) missions, in combination with CMB data. We show that the $f_{\\rm gas}$ experiment is likely to provide comparable constraining power to the best other, contemporary space and ground-based experiments described by the DETF. When combined, future CMB, SNIa, baryon acoustic oscillation (BAO), weak lensing, cluster number count and $f_{\\rm gas}$ experiments should provide precise, accurate constraints on $w(z)$ and allow significant progress in understanding the origin of cosmic acceleration. The structure of this paper is as follows: in Section~\\ref{sec:de} we define the dark energy model and the FoM. In Section~\\ref{sec:simdata} we describe the simulated $f_{\\rm gas}$ and CMB data sets. For the $f_{\\rm gas}$ data, we assume instrument characteristics appropriate for the baseline Constellation-X mission. The CMB data set approximates that expected from two years of Planck data. We also simulate a data set representative of that produced by follow-up observations of the Sunyaev-Zel'dovich effect in the clusters targeted for the $f_{\\rm gas}$ work. Section~\\ref{analysis} describes the Markov Chain Monte Carlo (MCMC) pipeline and details of the analysis method. Our main results are presented in Section~\\ref{constraints}. Section~\\ref{conclusions} summarizes our conclusions. ", "conclusions": "\\label{conclusions} \\begin{figure} \\includegraphics[width=3.2in]{Prospects_w0waP5_new.ps} \\caption{The 68 and 95 per cent confidence contours in the $w_{\\rm 0}-w_{\\rm a}$ plane determined from the $f_{\\rm gas}$+CMB data (black, dashed contours) using the default dark energy model and $5$ per cent systematic allowances. The solid red lines show the constraints obtained from the $f_{\\rm gas}$ data alone, using priors on $\\Omega_{\\rm b}h^2$, $\\Omega_{\\rm dm}h^2$, $l_{\\rm a}$ and $h$, as described in the text (Section~\\ref{cmbede}). The blue, dotted lines show the constraints from the $f_{\\rm gas}$ alone using priors on $\\Omega_{\\rm b}h^2$ and $h$ and assuming flatness. The figure shows how the CMB data contribute in constraining dark energy, especially at early times.} \\label{fig:cmbede} \\end{figure} We have examined the ability of a future X-ray observatory, with capabilities similar to those planned for Constellation-X, to constrain dark energy via the $f_{\\rm gas}$ experiment. We find that $f_{\\rm gas}$ measurements for a sample of 500 hot ($kT_{2500}\\gsim5$keV), X-ray bright, dynamically relaxed clusters, with a precision of $\\sim 5$ per cent, can be used to constrain dark energy with a FoM of $15-40$. These constraints are comparable to those predicted by the DETF \\citep{Albrecht:06} for other leading, planned (DETF Stage IV) dark energy experiments. We also find that, for the $f_{\\rm gas}$ experiment, the FoM can be boosted up by at least $\\sim 40$ per cent by selecting an optimal redshift distribution of suitable clusters on which to carry out the $f_{\\rm gas}$ observations. Interestingly, the optimal redshift distribution of $f_{\\rm gas}$ measurments appears to be shifted towards low redshifts. As discussed in the text, a future $f_{\\rm gas}$ experiment will need to be preceded by a large X-ray or SZ cluster survey that will find hot, X-ray luminous clusters out to high redshifts. A survey such as that planned with the Spectrum-RG/eROSITA mission should find several thousand of such clusters. Short `snapshot' follow-up observations of the clusters with a new, large X-ray observatory should be able to identify a sample of $\\sim 500$ suitable systems for $f_{\\rm gas}$ work. Attaining a precision of $\\sim 5$ per cent with individual $f_{\\rm gas}$ measurements should be straightforward for an observatory with characteristics similar to Constellation-X, requiring exposure times of $\\sim 20$ks on average. We note that the population of galaxy clusters in the redshift, temperature and X-ray luminosity range of interest has already been partially probed by the MACS survey \\citep{Ebeling:01}; Chandra observations of MACS clusters are used extensively in current $f_{\\rm gas}$ studies \\citep{Allen:04,LaRoque:06,Allen:07}. The low-level of X-ray flux contamination from point sources observed in MACS clusters also alleviates the requirements on the instrumental PSF for dark energy work via the $f_{\\rm gas}$ method. In determining the predicted dark energy constraints, we have employed the same MCMC method used to analyze current data. The MCMC method encapsulates all of the relevant degeneracies between parameters and allows one to easily and efficiently incorporate priors and allowances in the analysis. We have included an array of such systematic allowances, with tolerances ranging from optimistic to pessimistic. Our technique differs from the DETF \\citep{Albrecht:06}, who use a simpler Fisher matrix approach in the prediction of dark energy constraints. Despite these differences, we have endeavored to make our calculations of the FoM (Section 2) as comparable as possible. Benchmarking our results against those of the DETF for other, future `Stage IV' dark energy experiments i.e. large, long-term missions, we find that the $f_{\\rm gas}$ experiment should provide a comparable FoM to future ground-based SNIa (FoM=$8-22$), space-based SNIa (FoM=$19-27$), ground-based BAO (FoM=$5-55$), space-based BAO (FoM=$20-42$) and space-based cluster counting (FoM=$6-39$) experiments. Formally, the predicted FoM for the $f_{\\rm gas}$ experiment is comparable to `pessimistic' scenarios for weak lensing experiments discussed by \\cite{Albrecht:06}, although the value falls short of the most optimistic DETF weak lensing predictions. The tight constraints on $\\Omega_{\\rm m}$ and $\\Omega_{\\rm de}$ for the $f_{\\rm gas}$ experiment will be of importance when used in combination with other techniques. Interestingly, the `pivot point' for the $f_{\\rm gas}$ experiment lies between those of the SNIa and BAO/weak lensing/cluster number count experiments, offering excellent redshift coverage in attempts to pin down the evolution of dark energy. We conclude that the $f_{\\rm gas}$ experiment offers a powerful approach for dark energy work, which should be competitive with and complementary to the best other planned dark energy experiments." }, "0710/0710.2997_arXiv.txt": { "abstract": "{Gas-rich dwarf galaxies are probably the closest counterparts to primeval objects we can find in the local Universe, therefore it is interesting to study their evolution in different astrophysical contexts.} {We study the effects of interstellar clouds on the dynamical and chemical evolution of gas-rich dwarf galaxies. In particular, we focus on two model galaxies similar to IZw18 and NGC1569 in comparison to models in which a smooth initial distribution of gas is assumed.} {We use a 2-D hydrodynamical code coupled with a series of routines able to trace the chemical products of SNeII, SNeIa and intermediate-mass stars. Clouds are simulated by adding overdense regions in the computational grid, whose locations are chosen randomly and whose density profiles match observed ones. We consider both cloud complexes put at the beginning of the simulation and a mechanism for continuous cloud formation. The clouds are inherently dynamically coupled to the diffuse gas, and they experience heat conduction from a hot surrounding gas.} {Due to dynamical processes and thermal evaporation, the clouds survive only a few tens of Myr. Due to the additional cooling agent, the internal energy of cloudy models is typically reduced by 20 -- 40\\% compared with models of diffuse gas alone. The clouds delay the development of large-scale outflows by mass loading, therefore helping to retain a larger amount of gas inside the galaxy. However, especially in models with continuous creation of infalling clouds, their bullet effect can pierce the expanding supershell and create holes through which the superbubble can vent freshly produced metals. Moreover, assuming a pristine chemical composition for the clouds, their interaction with the superbubble dilutes the gas, reducing the metallicity. The resulting final metallicity is therefore generally lower (by $\\sim$ 0.2 -- 0.4 dex) than the one attained by diffuse models. } {} ", "introduction": "\\label{intro} Gas-rich dwarf galaxies are commonly classified into low surface-brightness dwarfs, called dwarf irregulars (dIrrs), and higher surface-brightness objects, usually called blue compact dwarf (BCD) galaxies. These classes of galaxies tend to have low metallicities, blue colors and complex and chaotic gas phases. A large fraction of these galaxies shows an ongoing star formation (SF) or at least hints that this process has been quenched in the recent past. In this case, these objects are commonly referred to as {\\it starburst} galaxies and their gas consumption timescales are much shorter than the Hubble time (Kennicutt \\cite{ken98}), making this a transient phase of their evolution. Owing to the energy released by stellar winds and supernovae (SNe), intense episodes of SF are also associated to the development of galactic winds or at least of large-scale outflows. The broad distinction between these two phenomena is the final fate of the outwards-directed flow of gas: galactic winds generally exceed the escape velocity while outflows do not, therefore they tend to recede towards the center of the galaxy. Clear signatures of outflows are present in NGC1705 (Hensler et al. \\cite{hen98}; Heckman et al. \\cite{hek01}), NGC1569 (Martin, Kobulnicky \\& Heckman \\cite{mkh02}), NGC3079 (Cecil et al. \\cite{cec01}), IZw18 (Martin \\cite{m96}), NGC3628 (Irwin \\& Sofue \\cite{is96}) among others. Perhaps the best examples of large-scale outflows driven by SN feedback are at large redshifts (Pettini et al. \\cite{pet98}; Pettini et al. \\cite{pet01}). Although it is not certain, in any of the above-mentioned objects, that the metals will definitely leave the parent galaxy, indirect hints of the ubiquity of galactic winds are given by the mass-metallicity relation (Tremonti et al. \\cite{tre04}; Dave\\'e, Finlator \\& Oppenheimer \\cite{dfo06}) and effective yields (Garnett \\cite{gar02}). From a theoretical point of view, the study of the evolution of gas-rich dwarf galaxies through numerical simulations has been performed by several authors in the recent past. The overall picture is that the occurrence of large-scale outflows is initially driven by the thermal pressure of a very hot, high pressurized gas and is favored by a flat distribution of the interstellar medium (ISM), which allows an easy vertical transport of material. However, since the transport of gas along the disk is very limited, outflows are not able to eject a significant fraction of the ISM, whereas the fraction of ejected metals can be very large (D'Ercole \\& Brighenti \\cite{db99}; MacLow \\& Ferrara \\cite{mf99}; Recchi, Matteucci \\& D'Ercole \\cite{rmd}, hereafter RMD). For NGC1569, Martin et al. (\\cite{mkh02}) derived a supersolar metal content in the galactic wind from X-ray spectra but also advocated mass-loading of it with the ISM. Most of these studies, however, have focused on flows in homogeneous media, neglecting the multiphase nature of the ISM, although several attempts to perform multiphase hydrodynamical simulations have been made in the past, particularly using the so-called {\\it chemodynamical} approach (Theis, Burkert \\& Hensler \\cite{tbh92}; Rieschick \\& Hensler \\cite{rh00}; Hensler, Theis \\& Gallagher \\cite{hen04}). The multiphase nature of the ISM, in particular its clumpiness, is observationally well established in dwarf galaxies (Cecil et al. \\cite{cec01}; Cannon et al. \\cite{cann05}; Leroy et al. \\cite{leroy06}) and it has a solid theoretical background with the seminal work of McKee \\& Ostriker (\\cite{mo77}). According to this model, the ISM is composed by a cold neutral phase (representing the cores of molecular clouds), confined by a warm medium (with temperatures of the order of 10$^4$ K) and these two phases (which are in pressure equilibrium) are embedded in a hot, diluted intercloud medium (HIM), continuously produced by SN explosions and stellar winds. Sufficiently dense clouds can pierce the HIM without being swept up, so they can become embedded therein (Vieser \\& Hensler \\cite{vh07b}). At the interface between clouds and HIM, condensation-evaporation processes establish the final fate of the cloud and its impact on the development of a galactic wind. In two previous papers, we have studied the dynamical and chemical evolution of model galaxies similar to IZw18 (Recchi et al. \\cite{rec04}, hereafter Paper I) and NGC1569 (Recchi et al. \\cite{rec06}, hereafter Paper II). The main results can be briefly summarized as follows: \\begin{itemize} \\item most of the analyzed models develop large-scale outflows. These outflows carry out of the galaxy mostly the chemical elements freshly produced during the most recent episodes of SF, with large escape fraction of metals with delayed production (like Fe and N). \\item Models with very short burst(s) of SF can cool and mix the newly formed metals in a very short timescale, whereas, when the SF is more complex, most of the metals are either directly ejected outside the galaxy through galactic winds or are confined in a too hot medium, therefore cannot contribute to the chemical enrichment of the warm ionized medium observed by emission lines from the \\hii gas. \\item Models with complex and long-lasting SF episodes reproduce the chemical composition and the abundance ratios of the above-mentioned galaxies much better than models with bursting SF. \\end{itemize} In this paper we simulate models with structural parameters similar to IZw18 and NGC1569. We increase arbitrarily the gas density of some specific regions of the computational grid, in order to create a ``cloudy'' phase, and we address the question how and to which extent a ``cloudy gas phase'' alters the former results. The clouds possess a specific density profile and can be either added at the beginning of the simulation or continuously created during the evolution of the model. We then analyze the differences between the dynamical and chemical evolution of these models with the ones presented in Paper I and Paper II. We point out that, at variance with the above-mentioned works, in this paper we will not specifically look for the best initial setups and the best assumptions in order to reproduce chemical and dynamical features of well-known objects. We will just stress the main variations produced by a clumpy initial setup. For this reason, we will also consider models which failed in Paper I and II at reproducing the observations of IZw18 and NGC1569. The paper is organized as follows: in Sect.~\\ref{cloud} we briefly recall the evolution of a cloud embedded in a hot medium; in Sect.~\\ref{model} we present the model and the adopted assumptions in the simulations. Results are presented in Sect.~\\ref{results_fix} (models with clouds fixed at the beginning of the simulation) and in Sect.~\\ref{results_inf} (continuous creation of clouds). Finally, a discussion and some conclusions are drawn in Sect.~\\ref{discussion}. ", "conclusions": "\\label{discussion} In this paper we have computed the chemical and dynamical evolution of model galaxies, with structural parameters similar to IZw18 and NGC1569, but in which a complex of clouds has been added, both perturbing the initial gaseous distribution and creating clouds, at a rate which equals the SF rate, and with infall velocity of 10 km s$^{-1}$ along the polar direction. The main focus of our work has been the comparison of these models with those presented in previous publications, in which similar setups but a smooth distribution of gas was considered. We have seen that the clouds are subject to a variety of disruptive phenomena like evaporation (when embedded in a hot medium), formation of shocks, development of thermal instabilities (in particular the Kelvin-Helmholtz instability) and expansion due to the larger pressure compared to the surrounding interstellar medium. The average lifetime of the clouds is therefore relatively short, depending on the cloud size (which is not constant in our simulations) but being of the order of a few tens of Myr. In spite of their transient nature, the clouds leave a significant imprint on the dynamical and chemical evolution of dwarf galaxies. The clouds, when they evaporate inside the superbubble, produce mass loading, increase the mean density of the cavity and, therefore, enhance the radiative losses (which are proportional to the square of the density). This results in a significant decrease of the total thermal energy (of the order of $\\sim$ 20 -- 40\\% compared to the diffuse models, depending on the assumptions), therefore less energy to drive the development of a large-scale outflow. On the other hand, the relative motion of supershell and clouds, in particular when the clouds infall motion is considered, can structure, pierce and create holes and fingers in the expanding supershell. These holes destroy the spherical symmetry initially present and favor the rushing out of the highly pressurized gas contained in the cavity. Therefore, in spite of the reduced thermal energy budget, the creation of large-scale outflows is not suppressed but, in most of the explored cases, only slightly delayed. Complex structures and fingers are indeed relatively common features in galaxies showing large-scale outflows like NGC1800 (Hunter \\cite{hun96}), NGC4214 (MacKenty et al. \\cite{mack00}) or NGC1705 (Heckman et al. \\cite{hek01}). The pressure inside the cavity is reduced compared to diffuse models, therefore in any case the total amount of ejected pristine gas is very small (smaller than in the models with smooth gas distribution) and, when averaging the size of the supershell in any direction, it turns out to be smaller than in diffuse models. But the piercing of the supershell can lead to an ejection efficiency of freshly produced metals as high as the one attained by diffuse models. This has, of course, important consequences on the chemical evolution of these objects. Since the differential winds are not suppressed, the diminished thermal energy of these models does not imply an increase of metals inside the galactic regions. On the other hand, the dilution effect of clouds plays a dominant role in determining the final metallicity of our model galaxies. Since the clouds have primordial chemical composition, their destruction and mixing with the surrounding medium reduces the total chemical composition without altering the abundance ratios. This produces a final metallicity $\\sim$ 0.2 -- 0.4 dex smaller than the corresponding diffuse models. We have examined the effect of a different choice of the IMF slope and of the nucleosynthetic set of yields (in particular for what concerns intermediate-mass stars). Flatter-than-Salpeter IMF slopes lead to an excessive production of energy, able to unbind most of the gas before the end of the simulation. On the other hand, in models with steeper IMF the development of large-scale outflows is almost completely suppressed. Different sets of intermediate-mass stars yields affect in particular the log(N/O) ratio. Renzini \\& Voli (\\cite{rv81}) yields tend to overestimate the primary production of nitrogen. When compared to the results of models implementing van den Hoek \\& Groenewegen (\\cite{vg97}) yields, the results differ by $\\sim$ 0.3 dex. Due to the assumption of a metallicity-dependent cooling function, also the dynamics is affected by the choice of the nucleosynthetic prescriptions. Our main results can be briefly summarized as follows: \\begin{itemize} \\item the clouds suffer thermal instabilities, formation of shocks and evaporation, therefore their lifetimes is limited to a few tens of Myr. \\item In spite of that, they are able to increase the main density of the cavity, provoking a reduction of the total thermal energy by $\\sim$ 20 -- 40\\% compared with a diffuse model. \\item The interaction clouds-supershell leads to strong structuring and piercing of the shell (in particular for models with continuous creation of infalling clouds), allowing the venting out of metals in spite of the reduced thermal energy. The development of large-scale outflows is therefore generally delayed but the ejection efficiency of metals remains unchanged. \\item From a chemical point of view, the effect of the clouds is to significantly reduce the total metallicity of the galaxies, without altering the abundance ratios. \\end{itemize}" }, "0710/0710.1285.txt": { "abstract": "We present results of long-slit spectroscopy in several positions of the Orion nebula. Our goal is to study the spatial distribution of a large number of nebular quantities, including line fluxes, physical conditions and ionic abundances at a spatial resolution of about 1$''$. In particular, we have compared the O$^{++}$ abundance determined from collisionally excited and recombination lines in 671 individual 1D spectra covering different morphological zones of the nebula. We find that protoplanetary disks (proplyds) show prominent spikes of {\\elect}([{\\nii}]) probably produced by collisional deexcitation due to the high electron densities found in these objects. Herbig-Haro objects show also relatively high {\\elect}([{\\nii}]) but probably produced by local heating due to shocks. We also find that the spatial distribution of pure recombination {\\oii} and {\\foiii} lines is fairly similar, in contrast to that observed in planetary nebulae. The abundance discrepancy factor (ADF) of O$^{++}$ remains rather constant along the slit positions, except in some particular small areas of the nebula where this quantity reaches somewhat higher values, in particular at the location of the most conspicuous Herbig-Haro objects: HH 202, HH 203, and HH 204. There is also an apparent slight increase of the ADF in the inner 40$''$ around $\\theta^1$ Ori C. We find a negative radial gradient of {\\elect}([{\\oiii}]) and {\\elect}([{\\nii}]) in the nebula based on the projected distance from $\\theta^1$ Ori C. We explore the behavior of the ADF of O$^{++}$ with respect to other nebular quantities, finding that it seems to increase very slightly with the electron temperature. Finally, we estimate the value of the mean-square electron temperature fluctuation, the so-called {\\ts} parameter. Our results indicate that the hypothetical thermal inhomogeneities --if they exist-- should be smaller than our spatial resolution element. ", "introduction": "\\footnotetext[1]{Based on observations made with the 4.2m William Herschel Telescope (WHT) operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof\\'\\i sica de Canarias.} \\label{intro} The analysis of the spectrum of {\\hii} regions allows to determine the chemical composition of the ionized gas phase of the interstellar medium from the Solar Neighbourhood to the high-redshift Universe. Therefore, it stands as an essential tool for our knowledge of the chemical evolution of the Universe. In photoionized nebulae, the abundance of the elements heavier than He is usually determined from collisional excitation lines (hereinafter CELs), whose intensity depends exponentially on the electron temperature, {\\elect}, of the gas. It was about 20 years ago when the first determinations of C$^{++}$ abundance from the intensity of the weak recombination line (hereinafter RL) of {\\cii} 4267 \\AA\\ were available for planetary nebulae (PNe). The comparison of the abundance obtained from {\\cii} 4267 \\AA\\ and from the CELs of this ion in the ultraviolet (UV) showed a difference that could be as large as a order of magnitude in some objects \\citep[e.g.][]{french83, rolastasinska94, mathisliu99}. \\citet{peimbertetal93} were the first in determinig the O$^{++}$ abundance from the very weak RLs, obtaining the same qualitative result: the abundances obtained from RLs are higher than those determined making use of CELs. This observational fact is currently known as the ``abundance discrepancy\" (hereinafter AD) problem. In the last years, our group has obtained a large dataset of intermediate and high resolution spectroscopy of Galactic and extragalactic {\\hii} regions using medium and large aperture telescopes \\citep{estebanetal02, estebanetal05, garciarojasetal04, garciarojasetal05, garciarojasetal06, garciarojasetal06b, lopezsanchezetal06}. The general result of these works is that the O$^{++}$/H$^+$ ratio calculated from RLs is between 0.10 and 0.35 dex higher than the value obtained from CELs in the same objects. The value of the AD that we usually find in {\\hii} regions is rather similar for all objects and ions and is much lower than the most extreme values found in PNe. The results for {\\hii} regions obtained by our group are fairly different that those found for PNe, and seem to be consistent with the predictions of the temperature fluctuations paradigm formulated by \\citet{peimbert67}, as it is argued in \\citet{garciarojas06} and \\citet{garciarojasesteban07}. In the presence of temperature fluctuations (parametrized by the mean square of the spatial variations of temperature, the so-called $t^{\\rm 2}$ parameter) the AD can be naturally explained because the different temperature dependence of the intensity of RLs and CELs. The existence and the origin of temperature fluctuations are still controversial problems and a challenge for our understanding of ionized nebulae. Recently, \\citet{tsamispequignot05} and \\citet{stasinskaetal07} have proposed an hypothesis to the origin of the AD, which is based on the presence of cold high-metallicity clumps of supernova ejecta still not mixed with the ambient gas of the {\\hii} regions. This cold gas would produce most of the emission of the RLs whereas the ambient gas of normal abundances would emit most of the intensity of CELs. Our group is interested in exploring on what variable or physical process the AD depends from different approaches. One of the most promising is based on the study of the behaviour of this magnitude at small spatial scales, something that has still not been explored in depth in nearby bright Galactic {\\hii} regions. In this paper, we make use of deep intermediate-resolution long-slit spectroscopy of the Orion nebula to study the dependence of the AD with respect to different nebular parameters: electron temperature and density, local ionization state of the gas, presence of high velocity material, and its correlation with different morphological structures (proplyds, ionization fronts, globules, Herbig-Haro objects), in {\\hii} regions. The spatial distribution of the physical conditions in the Orion nebula has been investigated by several authors. \\citet{baldwinetal91} obtained the density and temperature distribution in 21 and 14 points, respectively, along a 5$'$ line west of $\\theta^1$ Ori C, finding a density gradient that decreases to the outskirts of the nebula and a constant {\\elect}. \\citet{walteretal92} determined electron densities and temperatures and chemical abundances for 22 regions of the Orion nebula. Using also data from the literature, these authors find radial gradients of the physical conditions, but with a positive slope in the case of the temperature determined from {\\foiii} lines. \\citet{poggeetal92} obtained Fabry-Perot images of the inner 6$'$ of the nebula covering several bright CELs and taken with an average seeing of about 1\\farcs8. Those authors present a density map obtained from the ratio of the {\\fsii} doublet confirming the presence of a density gradient that reaches its highest point immediately south-southwest of the Trapezium stars, and some localized density enhancements in the Orion bar and some Herbig-Haro objects. Very recently, \\citet{sanchezetal07} have obtained an integral field spectroscopy mosaic of an area of 5$'\\times 6'$ of the center of the Orion nebula, with a spatial resolution of 2\\farcs7. The electron density map they obtain is consistent with that obtained by \\citet{poggeetal92} but richer in substructures, some of them possibly associated to Herbig-Haro objects. \\citet{sanchezetal07} also obtain a electron temperature map (derived from the line ratio of {\\fnii} lines) that shows clear spatial variations, which rise near the Trapezium and drop to the outer zones of the nebula. However, an important drawback of the temperature map of S\\'anchez et al. is that is based on non-flux calibrated spectra and possible effects due to variations in the dust extinction distribution cannot be disregarded. \\citet{odelletal03} obtained a high spatial resolution map of the electron temperature --derived from the line ratio of {\\foiii} lines-- of a 160$''\\times 160''$ field centered at the southwest of the Trapezium. The data were obtained from narrow-band images taken with the WFPC of the $HST$. Although they do not find a substantial radial gradient of {\\elect} in the nebula, \\citet{odelletal03} report the existence of small-scale temperature variations down to a few arcseconds compatible with the values of the temperature fluctuations parameter calculated from the AD determinations by \\citet{estebanetal04}. \\citet{rubinetal03} obtained $HST$/STIS long-slit spectroscopy at several slit positions on the Orion nebula analysing the electron temperature and density spatial profiles with resolution elements of 0\\farcs5 $\\times$ 0\\farcs5. These last authors do not find large-scale gradients of the physical conditions along the slits but a relatively large point-to-point variation and some correlation of such variations with several small-scale structures. The spatial mapping of the AD factor has been performed in few ionized nebulae but largely for PNe. \\citet{liuetal00}, \\citet{garnettdinerstein01}, and \\citet{krabbecopetti06} have found significant differences in the spatial profiles of the O$^{++}$/H$^+$ ratio derived making use of RLs and CELs suggesting the presence of chemical inhomogeneities or additional mechanisms for producing the {\\oii} lines in these objects. \\citet{tsamisetal03} have performed the only available study so far of the spatial distribution of the AD factor in an {\\hii} region: 30 Doradus. However, considering the extragalactic nature of this object and the spatial sampling of 3\\farcs5 used by those authors, their final spatial resolution is very low --about 1pc. In any case, \\citet{tsamisetal03} find a rather constant AD factor along the zone covered with their observations, a quite different behavior than that observed in PNe. In \\S\\S~\\ref{obsred} and~\\ref{linesel} of this paper we describe the observations, the data reduction procedure and the aperture extraction and measurement of the emission lines. In \\S~\\ref{phiscondabund} we derive the physical conditions and the ionic abundances from both kinds of lines: CELs and RLs. In \\S~\\ref{spat_prof} we present and discuss the spatial profiles of the physical conditions, line fluxes, and the abundance discrepancy factor along the slit positions. In \\S~\\ref{rad_dist} we discuss the large-scale radial distribution of some nebular properties along the nebula. In \\S~\\ref{cor_ADF} we explore possible correlations between the AD and different nebular parameters. In \\S~\\ref{t2} we address and estimate the possible temperature fluctuations inside the nebula. Finally, in \\S~\\ref{conclu} we summarize our main conclusions. ", "conclusions": "\\label{conclu} We have studied the spatial distribution of a large number of nebular quantities along five slit positions covering different morphological zones of the Orion nebula. The resolution element of the observations was 1\\farcs2 $\\times$ 1\\farcs03. The studied quantities were c(H$\\beta$), {\\elecd}, {\\elect}([{\\nii}]), {\\elect}([{\\oiii}]), the intensity of several selected lines (H$\\beta$, {\\cii} 4267 \\AA, {\\oii} 4649 \\AA, [{\\oiii}] 4959 \\AA, [{\\feiii}] 4881 \\AA, [{\\nii}] 5755 and 6584 \\AA, [{\\oi}] 6300 \\AA, and [{\\sii}] 6717 + 6731 \\AA), the O$^{++}$/H$^+$ ratio obtained from collisionally excited lines (CELs) and recombination lines (RLs), and the C$^{++}$/H$^+$ ratio obtained from RLs. The total number of apertures or 1D spectra extracted was 730. We have been able to determine the O$^{++}$/H$^+$ ratio from the faint RLs of this ion in a 92\\% of the apertures. The spatial distribution of {\\elecd} shows a large range of variation --larger than an order of mag\\-ni\\-tu\\-de-- across the nebula, with local maxima associated with the position of protoplanetary disks (proplyds), Herbig-Haro objects, the Orion bar, and the brightest area of the nebula at the southwest of the Trapezium. The proplyds show quite prominent spikes of {\\elect}([{\\nii}]) and much lesser ones of {\\elect}([{\\oiii}]). This fact could be due to collisional deexcitation on the nebular lines of {\\fnii} because of the high densities of these objects. Herbig-Haro objects also show somewhat higher values of {\\elect}([{\\nii}]) but, in this case, the origin could be related to extra heating of the gas due to shock excitation. The spatial distribution of the {\\oii} 4949 \\AA\\ and {\\foiii} 4959 \\AA\\ lines is fairly similar along all the slit positions, a very different behavior to that observed in planetary nebulae. We have found that the abundance discrepancy factor (ADF) of O$^{++}$ --the difference between the O$^{++}$ abundance determined from RLs and CELs-- remains, in general, rather constant along most of the observed areas of the nebula, showing values between 0.15 and 0.20 dex. However, there are some localized enhancements of the ADF, specially at the position of the Herbig-Haro objects HH 202, HH 203, and HH 204. The combined data of all slit positions indicate a clear decrease of {\\elect}([{\\nii}]) and {\\elect}([{\\oiii}]) with increasing distance from the main ionizing source of the nebula, $\\theta^1$ Ori C. On the other hand, the radial distribution of the ADF shows a rather constant value across the nebula except at the inner 40$''$, where the ADF seems to increase very slightly toward $\\theta^1$ Ori C. We have explored possible correlations between the ADF of O$^{++}$ and other nebular quantities, finding a possible very weak increase of the ADF for higher electron temperatures. There are not apparent trends between the ADF and c(H$\\beta$), {\\elecd}, the {\\elect}([{\\nii}])/{\\elect}([{\\oiii}]) ratio, O$^{++}$ abundance, and the O$^{++}$/O$^+$ ratio. Our spatially resolved spectroscopy allows to estimate the value of the mean-square electron temperature fluctuation in the plane of the sky, a lower limit to the traditional {\\ts} parameter. We find very low values in all cases, result that is in contradiction with previous estimates from the literature. Our results indicate that the hypothetical thermal inhomogeneities --if they exist-- should be lower than our spatial resolution limit of about 1$''$. It is clear that further studies on the {\\elect}, chemical abundances, and ADF distributions at sub-arcsec spatial scales are necessary in trying to disentangle (a) whether small spatial scale temperature fluctuations and/or metal-rich droplets are really present in the Orion nebula and HII regions in general, and (b) the origin of AD problem and its possible relation with {\\ts} and other nebular properties. The observations needed for this task are very difficult even for ground-based large-aperture telescopes and, by now, unfeasible with the current space telescopes and their available instrumentation. We thank G. Stasi\\'nska and M. Rodr\\'{\\i}guez for their fruitful comments and help. We are grateful to the referee, Y. Tsamis, for his careful reading of the paper and his comments. This work has been funded by the Spanish Ministerio de Ciencia y Tecnolog\u00eda (MCyT) under project AYA2004-07466." }, "0710/0710.2359_arXiv.txt": { "abstract": "{} {We evaluate the generation of magnetosonic waves in differentially rotating magnetized plasma.} {Differential rotation leads to an increase of the azimuthal field by winding up the poloidal field lines into the toroidal field lines. An amplification of weak seed perturbations is considered in this time-dependent background state.} {It is shown that seed perturbations can be amplified by several orders of magnitude in a differentially rotating flow. The only necessary condition for this amplification is the presence of a non-vanishing component of the magnetic field in the direction of the angular velocity gradient.} {} ", "introduction": "In astrophysical bodies, differential rotation is often associated with magnetic fields of various strength and geometry. If the poloidal field has a component parallel to the gradient of angular velocity, then differential rotation can stretch toroidal field lines from the poloidal ones. In the presence of the magnetic field, differential rotation can be the reason for various magnetohydrodynamic instabilities, particularly if the field geometry is complex. Some of these instabilities occur in the incompressible limit (Velikhov 1959; Fricke 1969; Acheson 1978; Balbus \\& Hawley 1991) which applies if the magnetic field is weak and the Alfv\\'en velocity is smaller than the sound speed. Other instabilities become important only in sufficiently strong magnetic fields when the effect of compressibility plays a significant role (Pessah \\& Psaltis 2005; Bonanno \\& Urpin 2006, 2007). Note that incompressible instabilities can often be suppressed by a strong magnetic field. For example, the magnetorotational instability (MRI) does not occur if the magnetic field satisfies the condition $B^2 > -8 \\pi \\rho s \\Omega \\Omega' L^2$ where $\\rho$ is the density, $L$ is the lengthscale of disturbances, $\\Omega=\\Omega(s)$ is the angular velocity, and $s$ is the cylindrical radius; $\\Omega'= d \\Omega/ds$ (see, e.g., Urpin 1996, Kitchatinov \\& R\\\"udiger 1997). Moreover, if $\\Omega$ increases with the cylindrical radius MRI cannot arise. In recent years, many simulations of differentially rotating magnetized bodies have been performed, and much of the dynamics was interpreted as being a direct consequence of the MRI (Brandenburg et al. 1995; Hawley at al. 1995; Matsumoto \\& Tajima 1995; Hawley 2000). Obviously, the MRI cannot be the only instability that operates in a rotating magnetized gas. For example, stratification can lead to a number of strong non-axisymmetric instabilities (Agol et al. 2001; Narayan et al. 2002; Keppens et al. 2002). Blokland et al. (2005) consider the influence of a toroidal field on the growth rate of the MRI and find that it leads to overstability (complex eigenvalue). Van der Swaluw et al. (2005) study the interplay between different instabilities and argue that the growth rate of convection can be essentially increased due to magnetorotational effects. Note, however, that these studies treat the stability of the magnetic field with a vanishing radial component, a condition which is often not met in astrophysical bodies. In fact, the presence of a radial magnetic field can change substantially the stability properties (Bonanno \\& Urpin 2006, 2007). In this paper, we consider stability of a differentially rotating gas in the presence of a non-vanishing radial magnetic field. Differential rotation causes the azimuthal field to increase with time by winding up the poloidal field lines into the toroidal ones. Therefore, a development of small perturbations occurs in the time-dependent background state. We show that stretching of the azimuthal field leads to the generation of magnetosonic waves in a flow. Magnetohydrodynamic waves and turbulence generated by this instability can play an important role in enhancing transport processes in various astrophysical bodies, such as accretion and protoplanetary Disks, galaxies, stellar radiative zones, etc. ", "conclusions": "We have shown that the winding up of toroidal field lines from the poloidal field lines is accompanied by an amplification of seed perturbations of the velocity and magnetic field. A very simplified model has been considered in this paper, but we believe that qualitatively the same results can be obtained for more general background states and perturbations. Differential rotation and compressibility of the gas lead to the generation of magnetosonic waves with the amplitude that grows with time. The physical processes responsible for this amplification are exactly the same that result in the instability considered by Bonanno \\& Urpin (2006, 2007). The only difference is that, in this paper, we consider the development of perturbations on a time-dependent background state and, as a result, the growth of perturbations is not exponential. The behaviour of seed perturbations depends essentially on various parameters and can generally be rather complex. If the parameter $q=s \\Omega'/\\Omega$ is relatively small ($\\leq 0.1$), then the perturbations of velocity and magnetic field initially grow monotonously and can reach quite high values. For example, the perturbation of the vertical velocity becomes approximately $150-200$ times greater than its initial value after only 15-30 rotation periods (see Figs.~2 and 3). Perturbations of the magnetic field reach even higher values during the initial stage. For instance, the Alfv\\'en velocity corresponding to the perturbation of the radial field component, $C_{As}$, can exceed the initial velocity perturbation by a factor $\\sim (1-3) \\times 10^5$ after the same time, but the perturbations of the toroidal field are even stronger. As a result of such a strong initial amplification, seed perturbations can already reach a non-linear regime after 15-30 rotation periods if their initial values are sufficiently large. Further evolution of perturbations will then be entirely determined by non-linear effects. However, if the non-linear regime is not reached during this initial stage, the behaviour of perturbations becomes oscillatory with slowly growing amplitude ($\\propto t^{1/2}$). At sufficiently large $t$, the frequency of oscillations grows linearly with time and is given approximately by \\begin{equation} \\omega \\approx \\Omega \\left( \\frac{1}{2} k H \\sqrt{\\beta_s} s \\Omega' t \\right). \\end{equation} Note that perturbations of the magnetic field exhibit more regular behaviour because they can be expressed in terms of the time integrals of a rapidly oscillating velocity. In the case of a strong differential rotation ($q \\sim 1$), perturbations exhibit the oscillatory behaviour from the very beginning and the initial growth of their amplitude is less significant. The generation of magnetosonic waves occurs even if the magnetic field is very strong and suppresses different MHD-instabilities which can arise in a differentially rotating flow (for example, the MRI). The presence of differential rotation and radial magnetic field is, however, crucially important for the considered process. Since both differential rotation and radial field are quite common in astrophysics, we believe that the considered mechanism can occur in various astrophysical bodies and plays an important role in enhancing transport processes in plasma. \\vspace{0.5cm} \\noindent {\\it Acknowledgments.} This research project has been supported by a Marie Curie Transfer of Knowledge Fellowship of the European Community's Sixth Framework Programme under contract number MTKD-CT-002995. VU thanks INAF-Ossevatorio Astrofisico di Catania for hospitality." }, "0710/0710.0878_arXiv.txt": { "abstract": "Using a population synthesis approach, we compute the total merger rate in the local Universe for double neutron stars, double black holes, and black hole -- neutron star binaries. These compact binaries are the prime source candidates for gravitational-wave detection by LIGO and VIRGO. We account for mergers originating {\\em both\\/} from field populations and from dense stellar clusters, where dynamical interactions can significantly enhance the production of double compact objects. For both populations we use the same treatment of stellar evolution. Our results indicate that the merger rates of double neutron stars and black hole -- neutron star binaries are strongly dominated by field populations, while merging black hole binaries are formed much more effectively in dense stellar clusters. The overall merger rate of double compact objects depends sensitively on the (largely unknown) initial mass fraction contained in dense clusters ($f_{\\rm cl}$). For $f_{\\rm cl} \\lesssim 0.0001$, the Advanced LIGO detection rate will be dominated by field populations of double neutron star mergers, with a small but significant number of detections $\\sim 20$ yr$^{-1}$. However for a higher mass fraction in clusters, $f_{\\rm cl} \\gtrsim 0.001$, the detection rate will be dominated by numerous mergers of double black holes originating from dense clusters, and it will be considerably higher, $\\sim 25 - 300$ yr$^{-1}$. In addition, we show that, once mergers of double black holes are detected, it is easy to differentiate between systems formed in the field and in dense clusters, since the chirp mass distributions are strikingly different. If significant field populations of double black hole mergers are detected, this will also place very strong constraints on common envelope evolution in massive binaries. Finally, we point out that there may exist a population of merging black hole binaries in intergalactic space. ", "introduction": "Gravitational wave astronomy is entering a new era: LIGO (Abramovici et al.\\ 1992) has now taken a full year of data at its design sensitivity; VIRGO (Bradaschia et al.\\ 1990) is nearing completion. Other detectors like GEO or TAMA are also on track. There are well defined plans for improving the LIGO and VIRGO detectors so that, within the next few years, their sensitivity will increase by a factor of up to 30. The list of potential sources for these high-frequency detectors is long and includes supernova explosions, neutron star oscillations, and persistent radiation from rapidly rotating, nonaxisymmetric neutron stars. However, compact object binaries remain the most promising sources. It has been known for some time that their formation may take place in two very different environments: in the galactic field, where their progenitors are massive binaries that evolve in isolation, and in dense star clusters, where they can form at high rates through dynamical interactions. The properties of the populations of double compact object binaries have been previously investigated using both observational and theoretical approaches. In the typical observational approach, the known compact objects binaries, i.e., radio pulsars in double neutron star (NS-NS) systems, were analyzed in detail. Based on their observed properties and the radio selection effects the properties of the entire population can be reconstructed and the expected merger rate can be calculated (Kim et al.\\ 2005). This approach, however, cannot be extended to black hole -- neutron star (BH-NS) binaries and double black hole (BH-BH) binaries since these systems have never been observed. Moreover, we know only one merging NS-NS binary in a globular cluster, and so this method has little power for the population originating in clusters. The second, theoretical, approach is based on numerical simulations of stellar evolution in field populations, combined with dynamical simulations of star clusters. For field populations some recent examples of theoretical studies include those by Belczynski, Kalogera \\& Bulik (2002, hereinafter BKB02), Voss \\& Tauris (2003), Pfahl et al.\\ (2005), Dewi et al.\\ (2006) and Belczynski et al.\\ (2007a). The importance of globular cluster evolution for double compact object formation has long been suspected. The fate of BH populations in globular clusters was studied by a number of groups (e.g., Sigurdsson \\& Hernquist 1993; Kulkarni, Hut \\& McMillan 1993; Portegies Zwart \\& McMillan 2000; Merritt et al.\\ 2004). However, previous studies have employed only very simplified treatments of stellar evolution for single stars and binaries in clusters, focusing instead mostly on dynamical interactions. It was pointed out (Phinney et al.\\ 1991; and more recently Grindlay, Portegies Zwart \\& McMillan 2006) that the formation of NS-NS and BH-NS binaries in clusters is not very efficient and the contribution of clusters to their total merger rate is rather small ($\\sim 10-30 \\%$). Although the predicted formation rate of merging double NS has a strong density dependence (Ivanova et al.\\ 2007), even for dense clusters like like 47~Tuc the NS contribution to total cluster merger rates should be overwhelmed by double BH mergers. Such systems are expected to be formed via dynamical interactions in cluster cores (Gultekin, Miller \\& Hamilton 2004; O'Leary et al.\\ 2006, 2007). O'Leary et al.\\ (2006, 2007) used realistic initial conditions for their cluster simulations (which included the stellar evolution of massive stars and binaries in the first few Myr of the cluster life) and assumed immediate creation of a BH subcluster. Their calculations did not consider further stellar or binary evolution and neglected the effects of lower-mass stars on BH populations. Their results provide the most up-to-date estimates of BH-BH merger rates from globular clusters and the corresponding LIGO detection rates assuming rapid BH segregation into an isolated subcluster. Mackey et al.\\ (2007) investigated the effects of BHs on the structural evolution of globular clusters. They used a realistic initial mass function for single stars and observed rapid mass segregation of the BHs into the cluster core (as expected from the Spitzer instability; see Watters et al.\\ 2000 and references therein). However, this study was based on direct $N$-body simulations that did not include {\\em any\\/} primordial binaries, even though binaries could affect the mass segregation and subsequent dynamics very significantly. In this paper we take the next step and calculate the evolution of representative star clusters from the onset of star formation taking into account the large binary fraction for massive stars. We assume that inelastic interactions of hard binaries in the cluster core are effective enough to prevent BHs from completely separating from the rest of the cluster, and we therefore treat the BHs as always well mixed and in thermal equilibrium with other stars in the cluster core. This simplifying assumption is opposite, and complementary, to the one adopted in the models of O'Leary et al. (2006, 2007). Their results and the merger rates calculated in this work give, respectively, lower and upper bounds on the LIGO detection rates. In our new models we include both stellar dynamics and full stellar evolution for single and binary stars to predict the merger rate of double compact objects. All stellar populations are evolved and allowed to interact through an entire cluster lifetime ($\\sim 13$ Gyr). Then we combine our estimates for star clusters and the merger rates calculated in O'Leary et al.\\ (2006) with the most recent population synthesis field calculations (Belczynski et al.\\ 2007a) to deduce the total merger rate. In Section~2 we present a description of the stellar binary evolution treatment used and we introduce our model for globular clusters. In Section~3 we present our results and in Section~4 a brief discussion. ", "conclusions": "\\label{summary} We have presented the ranges of total cosmic merger rate of double compact objects in the local Universe. Our results apply to the distances that will be reached by ground based gravitational-wave detectors such as LIGO or VIRGO ($\\sim 300$ Mpc for a typical NS-NS system). In calculating the rates we have considered formation of close double compact objects in field populations as well as in dense stellar clusters, i.e., globular clusters. The main result of our study shows that the predicted merger rates are too small for detection with the current instruments (i.e., initial LIGO) but are very promising for the upgraded detectors (i.e., advanced LIGO). We find that field populations dominate the formation of close NS-NS and BH-NS systems. This was already predicted by Phinney (1991) on the basis of observed binary pulsars. Recently Grindlay, Portegies Zwart \\& McMillan (2006) have shown that globular clusters can contribute only up to $\\sim 20 \\%$ to the cosmic NS-NS merger rates. Formation of double compact object systems with NSs is inhibited by the rather large natal kicks NSs receive in supernovae (Hobbs et al.\\ 2005). These lead to {\\em (i)} disruption of binaries hosting NS progenitors and {\\em (ii)} ejection of NSs from the cluster. The remaining systems containing a NS interact with heavier stars, and in particular BHs, and through exchange interactions are removed from binaries and do not form double compact objects. Also, the initial cluster mass that we use in our simulations ($\\sim 5 \\times 10^5\\msun$, which corresponds to the current average globular cluster mass in our Galaxy; e.g., Meylan \\& Heggie 1997, see their Fig.~10.2) is too small to produce a significant number of NS-NS or BH-NS systems. In particular, in our cluster simulations, we do not find any mergers of NS-NS nor BH-NS systems. The results for field populations of NS-NS and BH-NS mergers were discussed in detail by Belczynski et al.\\ (2007a). They find (their model A) that field Galactic merger rates are $\\sim 15$ Myr$^{-1}$ and $\\sim 0.1$ Myr$^{-1}$ for NS-NS and BH-NS, respectively. In our work we have adopted these rates as the total NS-NS and BH-NS merger rates (field + clusters) since the contribution from clusters is rather small for these systems. As discussed by Belczynski et al. (2007a) these rates are too small for any detection with the initial LIGO, while for advanced LIGO a small but significant number of detections is predicted: $\\sim 15$ yr$^{-1}$ and $\\sim 1$ yr$^{-1}$ for NS-NS and BH-NS, respectively. Formation of close BH-BH systems that merge in a Hubble time is expected to be very effective in clusters. This was already noted in studies that use realistic initial conditions for the evolution of BH-BH binaries in clusters (e.g., Miller \\& Hamilton 2002; Gultekin, Miller \\& Hamilton 2004; O'Leary et al.\\ 2006). Here, we have followed the evolution of a realistic average cluster with full stellar evolution and physical treatment of all BH-BH binaries, assuming that binary interactions are able to prevent the BHs from separating into an isolated subcluster. We have found that the production of BH-BH binary mergers under these assumptions is indeed remarkably effective in dense clusters. The most striking and counter-intuitive difference with other studies is that the predicted BH-BH merger rate in clusters is rather constant over long periods of time ($\\sim$ Hubble time). In earlier studies the BH populations were usually introduced in the cluster core and evolved separately from the other stars in the cluster. That leads to a larger merger rate at early times in the cluster evolution followed by a very rapid drop in the merger rates once BHs are eliminated (through mergers and ejections). However, the assumptions introduced in our work, namely {\\em (i)} stars that form BHs are initially placed throughout the cluster, and {\\em (ii)} BHs do not segregate so strongly as to form a completely decoupled subcluster and are instead allowed to interact with other stars in the cluster, make our results qualitatively different. We find that BHs that form in the core, in fact, produce mergers at the early stages of cluster evolution. But later many massive stars that were in the halo and that formed BHs will steadily feed the cluster core with BHs, i.e., continued mass segregation in the cluster halo is providing BHs to the cluster core on long timescales. Additionally, exchange interactions of BHs with unevolved (e.g., main sequence stars) in the core are effective over a long periods of time, and usually lead to formation of close BH-BH systems that merge later in the core. A simplified scenario involves two single BHs sinking into the core; each catches a main sequence companion, and then two BH-MS binaries interact, forming a close BH-BH system and two single main sequence stars are released back into the cluster. Our average cluster merger rate for BH-BH systems is $\\sim 3$ Gyr$^{-1}$ (see Figure~\\ref{f.merg}) for our simulated cluster with mass $M_{\\rm cl}=4.8 \\times 10^5 \\msun$. If we scaled this rate to the mass of the galactic disk ($M_{\\rm MW}=3.5 \\times 10^{10} \\msun$), then the BH-BH merger rate would come up to $\\sim 180$ Myr$^{-1}$ (i.e., this is the expected BH-BH merger rate if the entire mass of our Galactic disk were contained in dense globular clusters). We should compare this rate to the Galactic field BH-BH merger rate, $0.025$ Myr$^{-1}$ (Belczynski et al.\\ 2007a). It is clear that the production of BH-BH mergers is much more efficient (by $\\sim 3-4$ orders of magnitude) in clusters as compared to field evolution\\footnote{ We have used the rate from the same evolutionary model used for field population (model A of Belczynski et al.\\ 2007a) as it was employed in our standard cluster model.}. To combine our predicted cluster rates with the field rates we need to know the initial stellar mass fraction contained in clusters. If we look at the Galactic globular clusters we find that they contain about 0.001 of the total mass in stars found in the field ($f_{\\rm cl}=0.001$). However, it is reasonable to expect that many clusters may have been completely destroyed and that the clusters we see today were initially more massive and lost significant mass through evaporation (e.g., Vesperini 1998; Joshi et al.\\ 2001). Although we present our predictions for the entire range of plausible $f_{\\rm cl}$ values, we consider that the most reasonable estimate is still $f_{\\rm cl} \\gtrsim 0.001$. While we do not have reliable mass estimates for globular clusters in elliptical galaxies, it has been shown that the specific frequency of GCs per galaxy luminosity in ellipticals is significantly (about an order of magnitude) higher than in spiral galaxies (Kim \\& Fabbiano 2004). Moreover, elliptical galaxies are on average more massive than spiral galaxies. Therefore, an upper limit on the initial mass fraction contained in all clusters, although highly uncertain, can probably be set to $f_{\\rm cl} \\lesssim 0.01$. The total cosmic merger rate for BH-BH systems is then strongly dependent on the mass contained in globular clusters. For a small fraction ($f_{\\rm cl}=0.001$) we find the merger rate per Milky Way in our model to be $\\sim 0.2$ Myr$^{-1}$, while for a larger fraction ($f_{\\rm cl}=0.01$) the merger rate is $\\sim 2$ Myr$^{-1}$. At this point we can also estimate the detection rate for a given GW detector. We must keep in mind that the average chirp mass of field BH-BH binaries is much smaller ($M_{\\rm c} \\sim 7 \\msun$) than for cluster BH-BH mergers ($M_{\\rm c} \\sim 20 \\msun$). This allows us to observe cluster BH-BH mergers in a much larger volume and their relative contribution to the detection rate is greater than indicated simply by the merger rates (see \\S\\,~\\ref{ligo}). One could also imagine forming of few very massive BHs (up to $\\sim 100\\msun$) in the cluster (e.g., through mergers of massive binaries; Belczynski et al.\\ 2006) which could form BH-BH mergers characterized by extremely high chirp masses. Such mergers would be detectable from much greater distances, making the observed rates even higher. The predicted ranges of total detection rates, for all types of double compact objects, are presented in Figure~\\ref{f.rates} as a function of cluster contribution. The most likely range of values for this parameter ($f_{\\rm cl} \\sim 0.001\\div 0.01$) is marked with the vertical shaded area in Figure~\\ref{f.rates}. For very low cluster contributions ($f_{\\rm cl}=0.0001$) the detection rates correspond to mergers coming only from field populations and are adopted from the reference model of Belczynski et al.\\ (2007b; their model A). With increasing cluster contribution we see a drastic increase in predicted detection rates. This increase is connected to the very effective production of BH-BH mergers in clusters as discussed above. For advanced LIGO the detection rates could be as high as $\\sim 25-3000$ yr$^{-1}$ and are higher by more than an order of magnitude than the rates just for field populations. The total rates are dominated by dynamically formed BH-BH mergers in dense stellar clusters. If advanced LIGO does not observe this population of BH-BH mergers it will put strong constraints on the initial stellar mass fraction contained in dense stellar clusters. The production of BH-BH mergers in the field is inhibited by the process identified in Belczynski et al.\\ (2007b): many potential BH-BH progenitors evolve through a common envelope phase while the donor is evolving through the Hertzsprung gap. Such a common envelope leads most likely to a merger and aborts potential formation of a BH-BH system (because, in the Hertzsprung gap, the star has not yet developed a clear core-envelope structure and the inspiral does not stop before complete merger of the two interacting stars). If this current understanding of the common envelope phase is correct, we do not expect detection of more than a few field BH-BH mergers per year. However, if the progenitors somehow survive this phase, we could then expect up to $\\sim100$ detections of field BH-BH mergers per year by Advanced LIGO. Since the chirp mass distribution of the field and cluster populations are so different (see Figure~\\ref{f.mchirp} and Figure 4 of Belczynski et al.\\ 2007b) it would be easy to tell the two populations apart. This would, in turn, allow us to both {\\em (i)} derive the initial mass fraction in clusters, and {\\em (ii)} constrain the fate of massive binary systems going through a common envelope phase. This result highlights the importance of the chirp mass distribution as a diagnostic tool in gravitational wave astronomy (Bulik \\& Belczynski, 2003; Bulik, Belczynski, \\& Rudak, 2004)." }, "0710/0710.2750_arXiv.txt": { "abstract": "{The transition from galactic to extragalactic cosmic rays is discussed. One of critical indications for transition is given by the Standard Model of Galactic cosmic rays, according to which the maximum energy of acceleration for iron nuclei is of order of $E_{\\rm Fe}^{\\rm max} \\approx 1\\times 10^{17}$~eV. At $E > E_{\\rm Fe}^{\\rm max}$ the spectrum is predicted to be very steep and thus the Standard Model favours the transition at energy not much higher than $E_{\\rm Fe}^{\\rm max}$. As observations are concerned there are two signatures of transition: change of energy spectra and elongation rate (depth of shower maximum in the atmosphere $X_{\\rm max}$ as function of energy). Three models of transition are discussed: dip-based model, mixed composition model and ankle model. In the latter model the transition occurs at the observed spectral feature, ankle, which starts at $E_a \\approx 1\\times 10^{19}$~eV and is characterised by change of mass compostion from galactic iron to extragalactic protons. In the dip model the transition occures at the second knee observed at energy $(4 -8)\\times 10^{17}$~eV and is characterised by change of mass composition from galactic iron to extragalactic protons. The mixed composition model describes transition at $E \\sim 3\\times 10^{18}$~eV with mass composition changing from galactic iron to extragactic mixed composition of different nuclei. These models are confronted with observational data on spectra and elongation rates from different experiments, including Auger.} \\begin{document} ", "introduction": "The Ultra High Energy Cosmic Ray (UHECR) has two most important problems. One of them is a presence of spectrum features produced by propagation of UHECR particles through Cosmic Microwave Radiation (CMB) and the second is transition from galactic to extragalactic Cosmic Rays (CR). In the case of extragalactic protons two spectral signatures caused by interaction with CMB are predicted: Greisen-Zatsepin-Kuzmin (GZK) cutoff \\cite{GZK} and pair-production dip \\cite{BG88}. GZK cutoff is most spectacular prediction for UHECR, which status is still uncertain in present observations, though there are the indications to its presence. The pair-production dip is the spectral feature originated due to electron-positron pair production by extragalactic protons interacting with CMB: $p+\\gamma_{\\rm CMB} \\rightarrow p+e^++e^-$. Recently this feature has been studied in the works \\cite{Stanev2000,BGGPL,BGG}. The dip has been observed with very good statistical significance $\\chi^2$/d.o.f.$\\sim 1$ by the Fly's Eye, Yakutsk, Akeno-AGASA and HiRes detectors, and with much worse statistical significance by Auger detector. The pair-production dip and GZK cutoff are signatures of protons. The confirmation of the shape of these features is the evidence for proton-dominated composition of primary CRs. For nuclei as primaries the shape of the dip and GZK cutoff are strongly modified. The different explanation of the dip has been proposed by Hill and Schramm \\cite{HS85}. They interpreted the dip observed in 1980s in terms of two-component model. The low energy component can be either galactic or produced by Local Supercluster. The similar model has been considered in \\cite{YT}. The Hill-Schramm dip is widely used now for the explanation of the observed dip. From 1970s in the UHECR spectrum there was observed a flattening, which is called {\\em ankle}. Discovery of this feature at Haverah Park detector was interpreted as transition from the steep galactic component to more flat extragalactic one. The transition at ankle has been recently considered in \\cite{ankle}. In the dip model the transition is completed at the beginning of the dip at $E \\approx 1\\times 10^{18}$~eV. The ankle in this model appears as intrinsic part of the dip. Like in ankle model, the transition occurs here also as intersection of flat extragalactic component (this flatness is especially prominent in case of diffusive propagation) with steep galactic spectrum. In the dip and ankle models the extragalactic component is assumed to be proton dominated, while the galactic component is most probably composed by iron nuclei. In the {\\em intermediate model}, where transition occurs in the middle of the dip, the extragalactic CRs are assumed to have mixed composition \\cite{mixed}. In this paper all three above-mentioned models of transition are discussed. The logic of our discussion is as follows: we approach first the transition from the high energy end of galactic CRs, then we discuss the properties of UHECR relevant for transition problem and finally we describe the transition from properties of these two components. ", "conclusions": "\\vspace{1.3mm} The region of transition from galactic to extragalactic CRs at energy between $1\\times 10^{17} - 1\\times 10^{19}$~eV is the key energy range for understanding the origin of CRs. At low energy part it includes the high energy end of galactic CRs. The information on maximum energy of acceleration, chemical composition and propagation in Galaxy at these energies will clarify the total picture of origin at lower energies. The low energy part of UHECRs is important for understanding of origin of UHECRs and their propagation in extragalactic magnetic fields. The transition from galactic to extragalactic CRs is the central issue of this energy region. There are two detectors which cover partially the above-mentioned region: KASCADE-GRANDE \\cite{KASCADE-G} and TALE \\cite{TALE}. There are also the proposals to extend the observations of Auger to energy $E \\sim 1\\times 10^{17}$~eV (see e.g. \\cite{LE-Auger}). The Auger detector has great potential to explore this region, building more dense part of the detector covered with fluorescent, scintillator and muon detectors. The basic information which can be obtained includes precise measurement of energy spectra and mass composition (there is little hope to detect anisotropy in this energy region, though in some models the galactic sources can be observed in protons with energy $E \\lesssim 10^{18}$~eV \\cite{BGH}). At present we have the sufficiently good data on spectra and mass composition at energy range $1\\times 10^{18} - 4\\times 10^{19}$~eV. The spectra are measured with high statistics (especially in case of the Auger detector), but problem is the accuracy of energy determination. From quite disappointing Fig.~\\ref{fig:AgHiYaAu} (left panel) one concludes that scales of energy determination is quite different in all detectors. Energy calibration with help of the pair-production dip suggests that energy measured by scintillator detectors is systematically higher than that by the fluorescence detectors and it gives a reasonable recipe of increasing energies given by fluorescent method and decreasing it for the scintillation method. In this case the curves 'Yakutsk' and 'Akeno-AGASA' in Fig.~\\ref{fig:AgHiYaAu} go down and 'HiRes' and 'Auger' - up. For HiRes, AGASA and Yakutsk the method of calibration with help of dip works successfully (see Fig.~\\ref{fig:AgHiYa}) with energy shift within the allowed systematic errors, but for Auger it requires the shift by factor 2 greater than systematic error. The pair-production proton dip in terms of modification factor is an excellent tool to measure {\\em spectrum shape} independently of absolute flux. From Fig.~\\ref{fig:dips} one sees the excellent agreement of the theoretical dip with data of AGASA, HiRes and Yakutsk. By the standards of cosmic-ray physics the agreement with Auger data is also good, but $\\chi^2$ for comparison with SD data is very large. This is a result of very big statistics in the surface detectors at lowest energies $E \\geq 4.5\\times 10^{18}$~eV. In the lowest energy bin at $E=4.5\\times 10^{18}$~eV there are 4128 events and the error in determination of flux provided mostly by this statistics is $\\delta J/J=0.024$. The theoretical value of modification factor at this energy is only 14\\% higher than experimental value, but owing to very small $\\delta J/J$,the contribution of this bin to $\\chi^2$ is 99.27 ! Most probably the other sources of errors should be included in the bins with small $\\delta J/J$, and a possible source of this error is the energy errors which are changing randomly inside a bin. These could be statistical errors and energy-dependent part of systematic errors. Assuming that number of events are distributed in a bin as $N(E)=K E^{-\\gamma}$ one obtains $\\delta J/J = \\gamma (\\delta E/E)_r$, where $(\\delta E/E)_r$ is the random energy error inside the bin. The estimated value $\\delta J/J$ is much larger than what obtained in Auger analysis for all reasonable values of $(\\delta E/E)_r$ and $\\gamma$. More generally, according to Markus Roth's remark, $\\chi^2$ analysis is not adequate for the cases of small $\\delta J/J$ and large $(\\delta E/E)$. At this stage of analysis we do not consider Fig.~\\ref{fig:dips} as contradiction with Auger data.\\\\*[4mm] Coming to the transition from galactic to extragalactic CRs, we emphasize that at present there are only two experimental methods to study it: measuring the spectrum and mass composition. The transition will be clearly seen if spectrum of iron nuclei and that of protons are measured separately (see Fig.~\\ref{fig:dip-ankle}), but even without this ideal possibility the total spectrum has signatures of transition in the form of the spectral features - {\\em second knee} in case of the dip model and {\\em ankle} in case of the ankle model. The spectrum can be measured nowadays with high accuracy and its shape contains the information about mass composition, which is the other characteristic of the transition. The pair-production dip with its specific shape is a signature of proton-dominated composition (nuclei contribution should be not more than 10 -15 \\% \\cite{BGGPL}) and its observational confirmation is an argument not weaker than that due to $X_{\\rm max}$ measurement (we remind that only two free parameters are involved in describing about 20 energy bins in each experiment). \\\\*[4.1mm] The mass composition gives another way to test the transition. The best method at present is given by measuring of elongation rate $X_{\\rm max}(E)$. Unfortunately this method has many uncertainties, including those in value of fluorescent yield, absorption of UV light in the atmosphere and uncertainties in the models of interactions, needed to convert the tested mass composition into $X_{\\rm max}$. The systematic errors in measuring $X_{\\rm max}$ can be as large 30~g/cm$^2$ to be compared with difference about 100~g/cm$^2$ between $X_{\\rm max}$ for protons and iron. The better sensitivity for distinguishing different nuclei is given by distribution over $X_{\\rm max}$ \\cite{ABBO}. There are three models of the transition: ankle, dip and mixed-composition model. They differ most notably by the energy of transition (ankle: $E \\sim 1\\times 10^{19}$~eV, dip: $E \\approx 1\\times 10^{18}$~eV and mixed composition model $E \\approx 3\\times 10^{18}$~eV), and by mass composition of extragalactic component (protons - for the ankle model, proton-dominated - for the dip model and mixed composition - for the third model). The {\\em ankle model} contradicts the Standard Model of Galactic CRs (energy where galactic flux is half of that observed is two orders of magnitude higher than energy of iron knee) and severely disagrees with $X_{\\rm max}$ measured in all experiments at $(1.5 - 5)\\times 10^{18}$~eV. The {\\em dip model} is based on well confirmed signature of proton interaction with CMB - pair-production dip. The two other models must assume that agreement of pair-production dip with data is accidental and the observed dip is produced by two components, galactic and extragalactic. The dip model assumes the iron-dominated galactic flux below $5\\times 10^{17}$~eV and proton-dominated extragalactic flux above $1\\times 10^{18}$~eV. This mass composition is confirmed by HiRes and HiRes-Mia data for elongation rate. It does not contradict the bulk of all data on $X_{\\rm max}$, but contradicts $X_{\\rm max}$ measured by Auger, especially the highest energy points. The generation spectrum in this model is $E^{-2}$ or $E^{-2.2}$ as needed by shock acceleration with a steepening to $\\gamma_g=2.7$ due to distribution of sources over maximum energy of acceleration of source luminosities. The proton-dominated composition can be produced in some models of injection to the shock acceleration. The {\\em mixed composition model} assumes mixed composition generation spectrum for extragalactic component with generation index 2.1 - 2.3. It has many free parameters, most notably ones describing the mass composition of the generation spectrum, and thus it can in principle explain any observed mass composition. However, this model has a robust prediction at energy $E \\gtrsim 3\\times 10^{19}$~eV: proton-dominated composition and the GZK feature. As far as Auger elongation rate is concerned, the mixed composition model explains well the break in elongation rate at $2\\times 10^{18}$~eV and contradicts the two Auger points at $E > 2\\times 10^{19}$~eV. The energy where transition to extragalactic CRs is completed in most versions of this model equals $E \\approx 3\\times 10^{18}$~eV. Much better quality of data on $X_{\\rm max}$ is needed to distinguish the dip and mixed-composition models by $X_{\\rm max}$ measurements. Probably it is possible to do using $X_{\\rm max}$ distribution \\cite{ABBO}.\\\\*[1mm] We will comment now on agreement of the transition models with the measured galactic spectrum. For all three models it is reached by the formal subtraction procedure: the galactic spectrum is found as difference between measured total spectrum and calculated extragalactic spectrum. But the galactic spectrum calculated in the Standard Model at $E \\gtrsim 1\\times 10^{17}$~eV is very steep and, as was demonstrated in \\cite{Tanco}, for diffusive model of propagation all three models contradict the calculated galactic spectrum, the dip model to the less extent. Strictly speaking this contradiction is produced by exponential cutoff in the acceleration spectrum at $E > E_{\\rm max}^{\\rm acc}$. \\\\*[2mm] The most consistent conclusions on nature of observed UHECRs are obtained at present by HiRes detector: it has confirmed the pair-production dip and thus proton-dominant composition at $1\\times 10^{18} - 4\\times 10^{19}$~eV, the $X_{\\rm max}$ measurements agree with proton-dominant composition at $E > 1\\times 10^{18}$~eV , and $E_{1/2}$ measurement confirms that steepening of the spectrum observed at $E > 4\\times 10^{19}$~eV is really the GZK cutoff. Therefore, according to these data CRs observed at $E \\gtrsim 1\\times 10^{18}$~eV are extragalactic protons exhibiting two signatures of interaction with CMB: pair-production dip and GZK feature." }, "0710/0710.5036_arXiv.txt": { "abstract": "We study the 37 brightest radio sources in the Subaru/\\textit{XMM-Newton} Deep Field (SXDF). Using mid-IR (Spitzer MIPS 24 $\\mu \\rm m$) data we expect to trace nuclear accretion activity, even if it is obscured at optical wavelengths, unless the obscuring column is extreme. Our results suggest that above the `FRI/FRII' radio luminosity break most of the radio sources are associated with objects that have excess mid-IR emission, only some of which are broad-line objects, although there is one clear low-accretion-rate FRI. The fraction of objects with mid-IR excess drops dramatically below the FRI/FRII break, although there exists at least one high-accretion-rate QSO. Investigation of mid-IR and blue excesses shows that they are correlated as predicted by a model in which a torus of dust absorbs $\\sim$30\\% of the light, and the dust above and below the torus scatters $\\gtsim$1\\% of the light. ", "introduction": "Powerful radio sources are believed to have central super-massive black holes (SMBH) with uniformly high accretion rates at the highest radio luminosities and typically lower accretion rates at lower radio luminosities (\\cite{rs91}). Low-luminosity radio jets can, however, be associated with high-accretion-rate systems, and these so-called `radio quiet' quasars appear to have similar FRI-like radio structures to low-accretion-rate counterparts of similar radio luminosity (e.g. \\cite{hbr07}). At low redshift, the most massive ($\\gtsim$ $10^{8} \\rm M_{\\odot}$) SMBH typically have very low accretion rates with systematically higher average values at $z \\gtsim$ 2, the so-called `quasar epoch' (\\cite{yt02}). These observational results fit in with theoretical ideas that a `quasar mode' of feedback is prevalent in the distant universe, and that a `radio mode' feedback is dominant at low redshift (e.g. \\cite{cro06}). The central region of an AGN is surrounded by a dusty torus which absorbs light and re-emits it in the infrared. Above and below the plane of the torus, dust scatters light yielding a blue excess. Such mechanisms make it difficult to observe objects viewed through the torus directly in the optical, UV and soft X-rays. The torus creates anisotropic obscuration of the central regions resulting in two different types of observed objects, type 1 that are viewed face-on and type 2 that are viewed edge-on. Here we use mid-IR observations to search for evidence of accretion in a manner which is far less dependent on orientation. The sample studied here is the 37 brightest radio sources from the VLA survey of the Subaru/\\textit{XMM-Newton} Deep Field (SXDF; \\cite{sim06}) with flux densities greater than 2 mJy at 1.4 GHz. Optical, X-ray and radio observations of the SXDF were made within the 1.3 square degree Subaru/\\textit{XMM-Newton} Deep Field with Subaru, \\textit{XMM-Newton} and the VLA respectively. Thirteen of our objects are not as yet spectroscopically confirmed so we use photometric redshifts in these cases, calculated with the HYPERz code (\\cite{boz00}) and typically nine data points from $B-$band (440 nm) to 4.5 $\\mu \\rm m$ (Vardoulaki et al. in prep). ", "conclusions": "\\begin{figure} \\begin{center} \\setlength{\\unitlength}{1mm} \\begin{picture}(140,38) \\put(65,-21){\\special {psfile=\"l1_4_D.ps\" vscale=32 hscale=32 angle=90}} \\put(143,-21){\\special {psfile=\"blue_red.ps\" vscale=32 hscale=32 angle=90}} \\end{picture} \\end{center} \\vspace{0.5in} {\\caption[junk]{\\label{fig1} a {\\it Left}: Radio Luminosity $\\log_{10}(L_{1.4{\\rm GHz}}/ \\rm [W Hz^{-1} sr^{-1}])$ versus largest projected linear size $D$: symbols indicate optical/IR classification; filled red circles for quasars `Q'; filled blue triangles for obscured quasars `OQ'; filled black squares for possible galaxies `G?'; black squares for secure galaxies `G'; green upside-down triangles for starbursts 'SB'; orange diamonds for weak quasars `WQ'; and light blue stars for BL Lac `BL'. The horizontal lines show the RLF and FRI/FRII breaks calculated from the values in \\cite{fr74} using a typical steep-spectrum spectral index of 0.8 and translated to our assumed cosmology. b {\\it Right}: Blueness versus mid-IR excess. Symbols are the same as in the left figure. The black dotted line corresponds to the best-fit line in the log-linear plane where all objects, were treated as detections; the slope and intercept are 0.26 and -1.27 respectively, giving (see Eqn (2)) $\\rm k_{1} = 0.05$ and $\\rm k_{2} = 0.03$. The blue solid line corresponds to the best-fit line in log-linear plane where objects without detections at 24 $\\mu \\rm m$ were treated as limits; the slope and intercept are 0.10 and -0.41 respectively, giving $\\rm k_{1} = 0.39$ and $\\rm k_{2} = 0.09$. The Buckley-James method in the ASURV statistics package (\\cite{lav92}) was used in these calculations. `SB' and `BL' objects were excluded from the calculations since they have SEDs dominated by different physical processes to those assumed in the model described by Eqns (1) and (2). We adopt a radio spectral index $\\alpha$ = 0.8 ($S_{\\nu} \\propto \\nu^{-\\alpha}$), unless a spectral index could be calculated using 1.4 GHz data from \\cite{sim06} and 325 MHz data from \\cite{tasse} (see Vardoulaki {\\it et al.} in prep.). We assume throughout a low-density, $\\Lambda$-dominated Universe in which $H_{0}=70~ {\\rm km~s^{-1}Mpc^{-1}}$, $\\Omega_{\\rm M}=0.3$ and $\\Omega_{\\Lambda}=0.7$. }} \\end{figure} \\addtocounter{figure}{0} We use optical/IR observations to classify a radio source as either Quasar `Q', Obscured Quasar `OQ', Galaxy? `G?', Galaxy `G', Starburst `SB', Weak Quasar `WQ' or BL Lac `BL'. We deem that nuclear accretion is `significant' in objects that obey $\\log_{10}(L_{24 \\mu \\rm m}/ \\rm [W Hz^{-1} sr^{-1}]) > 23.1$ (or $[\\lambda L]_{24 \\mu \\rm m} > 10^{37.3}$ ${\\rm W}$). This value corresponds to $[\\lambda L]_{24 \\mu \\rm m} \\ge 10^{-1.8} L_{Edd}$, a typical lower limit for quasars (\\cite{mcl04}), for a black hole mass $M_{\\rm BH} \\ge 10^{8} M_{\\odot}$, a typical lower limit for radio sources (\\cite{mcl_etal04}); $L_{Edd}$ is the Eddington luminosity. We then define the following categories:\\\\ i) {\\bf Q}: Broad lines in the optical spectrum (3/37 cases). None of these are detected at 24 $\\mu \\rm m$, although their limits are insufficient to rule out significant accretion.\\\\ ii) {\\bf OQ}: Objects with a 24-$\\mu \\rm m$ detection (5/37 cases) and with sufficient $L_{24}$ to represent significant accretion. This class may be incomplete in that some objects in the `G?' class, as described next, have limits above this critical value.\\\\ iii) {\\bf G?}: A galaxy that has a 24-$\\mu \\rm m$ limit consistent with it lying above the $\\log_{10}(L_{24 \\mu \\rm m}/$ $\\rm [W Hz^{-1} sr^{-1}]) = 23.1$ line\\footnote{Because of the 24 $\\mu \\rm m$ flux density limit, these objects are at high redshift, and hence, because of the 1.4-GHz flux density limit, a high-$L_{1.4 \\rm GHz}$ sub-set of the objects lacking Spitzer 24 $\\mu \\rm m$ detections.} (11/37 cases).\\\\ iv) {\\bf G}: All other objects (15/37 cases) without significant accretion, unless they fall into three special categories defined by properties derived from spectroscopy, the SED and the optical structure: {\\bf SB}: evidence from the SED of a starburst component (1/37 cases); {\\bf WQ}: evidence from the SED of a quasar component but no 24 $\\mu \\rm m$ detection (1/37 cases); {\\bf BL}: featureless red continuum and a point source at $K$ (1/37 cases).\\\\ Figure 1a shows the 1.4-GHz radio luminosity at $L_{\\rm 1.4 GHz}$ versus the projected linear size $D$ with the symbols denoting the different optical/IR classes. We see that nearly all `Q', `OQ' and `G?' objects of our sample lie above the `FRI/FRII' luminosity break\\footnote{Although the FRI/FRII classification scheme is on the basis of radio structure, there is a sharp change in radio structure at a characteristic radio luminosity \\cite{fr74}.}, with the exception of the `OQ' sxdf\\_0034 (the `G?' object near sxdf\\_0034 lies very close to the boundary of significant accretion). In previous studies, the quasar fraction has been defined as the number of sources with quasar-like optical features (e.g. broad lines) and has a value of $\\sim 0.1 \\rightarrow 0.4$ over this range of $L_{1.4 \\rm GHz}$ (e.g. \\cite{wil00}). We introduce the `quasar-mode fraction' $f_{\\rm QM}$ to describe the fraction of objects with high accretion rates to the total number of objects. Above the FRI/FRII break $f_{\\rm QM} \\sim 0.5 - 0.9$ (the lower value assumes the 24 $\\mu \\rm m$ limits are much higher than the true 24 $\\mu \\rm m$ values, whereas the higher value assumes the true values lie just below the limits). The one clear exception in this regime is sxdf\\_0001, which has no evidence of a QSO and a clear Twin-Jet (FRI) radio structure. The quasar-mode fraction drops dramatically below the FRI/FRII break\\footnote{We note that objects in our sample above the FRI/FRII break have median redshift $z_{\\rm med} \\sim 1.6$, whereas those below have $z_{\\rm med} \\sim 0.65$, so evolutionary effects may also be important.}, and whether or not one excludes some of the compact ($D <$ 100 kpc) sources as potentially part of a separate (beamed) population, then $f_{\\rm QM} \\ltsim$ 0.1 because nearly all objects are galaxies `G'. The counter example here are sxdf\\_0034, the only `OQ' below the FRI/FRII break, and potentially an optically-obscured example of unobscured FRI QSOs already studied in this radio luminosity regime (e.g. \\cite{hbr07}). Inspection of the SEDs (Vardoulaki {\\it et al.} in prep) shows that some of our objects have an excess at 24 $\\mu \\rm m$ above that expected from extrapolation of the stellar populations. This is quantified via a measure of the mid-IR excess, $\\log_{10}([\\nu L]_{10 \\mu \\rm m rest} / [\\nu L]_{1\\mu \\rm m rest})$. A comparison of mid-IR excess and blueness is presented in Fig. 1b where a positive correlation is obvious. The generalised Spearman correlation calculated using survival analysis statistical package ASURV (\\cite{lav92}) is 0.657 with a 99\\% probability for a correlation. Consider a simple model in which blueness is connected to mid-IR excess through the following equations\\footnote{Equation (2) relies on the $ln(1+x) \\approx x$ approximation which is only accurate around and below the knees of the functions plotted in Fig. 1b.}: \\begin{equation} [\\nu L]_{4000 \\rm \\AA \\rm rest} = \\rm k_{1} \\times [\\nu L]_{1\\mu \\rm m rest} + \\rm k_{2} \\times [\\nu L]_{10 \\mu \\rm m rest} \\Rightarrow \\end{equation} \\begin{equation} \\log_{10}\\left(\\frac{[\\nu L]_{4000 \\rm \\AA rest}}{[\\nu L]_{1\\mu \\rm m rest}} \\right) = \\log_{10}(\\rm e) \\times \\frac{\\rm k_{2}}{\\rm k_{1}} \\times \\left(\\frac{[\\nu L]_{10 \\mu \\rm m rest}}{[\\nu L]_{1\\mu \\rm m rest}}\\right) + \\log_{10}(\\rm k_{1}), \\end{equation} where $\\rm k_{1}$ encodes the contribution of the stellar population of a passively evolving galaxy formed at high redshift ($z > 5$), and $\\rm k_{2}$ the mid-IR-excess parameter that we are looking to calculate for this sample of radio sources. This model assumes that light from the nucleus with intrinsic optical luminosity $L_{\\rm opt}$ is i) absorbed by dust and re-emitted in the mid-IR generating luminosity $[\\nu L]_{10 \\mu \\rm m rest}$ and ii) scattered, generating luminosity $[\\nu L]_{4000 \\rm \\AA rest}$. Fig. 1b shows best-fit lines for two scenarios: 1) all objects were treated as detections (black dotted line), and 2) objects are treated as upper limits according to their 24 $\\mu \\rm m$ detection (blue solid line), where in both cases `SB' and `BL' objects were excluded (Fig. 1b). Averaging these results we deduce $\\rm k_{1} \\sim 0.2$ and $\\rm k_{2} \\sim 0.05$, which agrees well with independent evidence. The value deduced for $\\rm k_{1}$ is in line with the expectations of template spectra of galaxies which formed their stars at high redshift. Optical polarisation studies (e.g. \\cite{kis01}) tell us that $[\\nu L]_{4000 \\rm \\AA rest} \\gtsim$ $0.01 [\\nu L]_{\\rm opt}$, which is consistent with our value of $\\rm k_{2}$ given that QSO SED studies suggest $[\\nu L]_{10 \\mu \\rm m rest} \\sim 0.3 [\\nu L]_{\\rm opt}$ (\\cite{rr95}). We conclude that whenever nuclear accretion is significant in our sample of radio sources, dust in the torus absorbs 30\\% of the photons and dust above and below the torus scatters $\\gtsim$1\\% of the photons." }, "0710/0710.0339_arXiv.txt": { "abstract": "The observational evidence for central black holes in globular clusters has been argued extensively, and their existence has important consequences for both the formation and evolution of the cluster. Most of the evidence comes from dynamical arguments, but the interpretation is difficult, given the short relaxation times and old ages of the clusters. One of the most robust signatures for the existence of a black hole is radio and/or X-ray emission. We observed three globular clusters, NGC6093 (M80), NGC6266 (M62), and NGC7078 (M15), with the VLA in the A and C configuration with a 3-$\\sigma$ noise of 36, 36 and 25 $\\mu$Jy, respectively. We find no statistically-significant evidence for radio emission from the central region for any of the three clusters. NGC6266 shows a 2-$\\sigma$ detection. It is difficult to infer a mass from these upper limits due to uncertainty about the central gas density, accretion rate, and accretion model. ", "introduction": "Although we do not understand how the nuclei of galaxies form or why they have black holes (BH) at their centers, the correlation between BH mass and bulge velocity dispersion does shed light on their formation and evolutionary histories (Gebhardt et al. 2000a, 2000b: Ferrarese and Merritt 2000). A number of different theories (e.g., Silk \\& Rees 1998; Haehnelt \\& Kauffmann 2000; Robertson et al. 2006) predict a BH mass bulge-velocity-dispersion relation, although they predict different slopes and intercepts for this relation. Exploration of the extreme ends of this relationship will help illuminate the underlying physical model, and in this paper we focus on the low mass end. Black holes at the low end of the relations, with masses between 100 and $10^6~\\Msun$, are generally referred to as intermediate-mass black holes (IMBHs). There is significant evidence that black hole masses from $10^5-10^6~\\Msun$ exist from the work of Barth, Greene \\& Ho (2005) and Greene \\& Ho (2006). To go to yet smaller black hole masses, an extrapolation of the correlation between black hole mass and stellar velocity dispersion suggests studying stellar systems with velocity dispersions of 10--20~\\kms. These dispersions are characteristic of globular clusters. Whether the existence of black holes in globular clusters could shed light on the formation and correlations of supermassive black holes is unknown, but clearly it is a possibility. Furthermore, the existence of massive black holes in clusters will have a significant effect on the cluster evolution. Thus, quantifying the demographics of black holes in clusters may be related to how supermassive black holes grow, and will definitely yield useful information about the evolution of clusters. Theoretical work suggests that we might expect IMBHs at the centers of steller systems (Ebisuzaki et al. 2001; Portegies Zwart \\& McMillian 2002; Miller \\& Hamilton 2002), although it appears to be difficult to make black holes more massive than 100~$\\Msun$. Gurkan et al. (2004) suggest that IMBHs may be easy to form through runaway collisions with massive stars. Discoveries of BHs in globular clusters have been claimed --- G1 in M31 (Gebhardt, Rich \\& Ho 2002) and M15 (van der Marel et al. 2002; Gerssen et al. 2002). In fact, the M15 claim has been made for the past 30 years, starting with the result of Newell, da Costa \\& Norris (1976) and subsequently challenged by Illingworth \\& King (1977). The basic issue is being able to distinguish a rise in the central mass-to-light ratio being due to either a black hole or the expected stellar remnants (neutron stars, massive white dwarfs and solar mass black holes). The most recent M15 result has been challenged by Baumgardt et al. (2003a). The result in G1 has also been challenged by Baumgardt et al. (2003b) but Gebhardt, Rich \\& Ho (2005) include additional data and analysis that support the black hole interpretation. There has been two further observations which strongly support the existence of a black hole in G1. Trudolyubov \\& Priedhorsky (2004) measure X-rays from G1 using the Chandra Observatory, centered to within 2\\arcsec\\ of the center of G1. Subsequently, Pooley \\& Rappaport (2006) suggest the X-ray emmission is from accretion onto a black hole, and Maccarone \\& Koerding (2006) point out that if a black hole is present then a 30 $\\mu$Jy radio source may be expected. The most significant observation comes from Ulvestad, Greene \\& Ho (2007) who find a 28 $\\mu$Jy (4.5$\\sigma$) emission centered on G1. Other interpretations are a pulsar wind or a planetary nebula. The pulsar wind seems unlikely given the age of G1 and the point-like radio source (an old pulsar would have a large size). A planetary nebula would show optical emission lines which are not seen in the HST or Keck spectra of Gebhardt et al. (2003). Other studies of the existence of black holes in globular clusters have been less compelling. Colpi, Mapelli, \\& Possenti (2003) use indirect dynamical arguments to suggest a few hundred solar mass black hole in NGC~6752. McLaughlin et al. (2006) provide an estimate of black hole in 47Tuc of $900\\pm900~\\Msun$. To date, there are no published upper limits of black hole masses that are significantly below that expected from an extrapolation of the correlation between black hole mass and stellar velocity dispersion. While the dynamical arguements strongly support the black hole interpretation in at least G1, the radio emission provides a clear and obvious result. Unfortunately, it is difficult to predict the radio emission from a given black hole mass. The next step is to explore other globular clusters with a similar setup and deep exposures. ", "conclusions": "Failure to detect radio radiation at 8.6 GHz from the centers of three globular clusters does not prove that no globular clusters have IMBHs at their centers. Besides not having a black hole, other interpretations include 1) accretion by the BH could be episodic and we happened to observe the BHs in an ``off-state\", 2) the gas density could be much lower compared to galaxies, 3) the radiative efficiency may be lower than assumed (although the assumed efficiencies are already quite low), 4) or the accretion model may not be adequate in general. We would predict, using the relation of Merloni et al. (2003) or using standard accretion models and gas density estimates (as done in Maccarone 2004), that we should have detected radio radiation at 8.6 GHz if accretion is steady and the accretion rate times the Bondi rate is 10$^{-4}\\times$ or higher. We would not have been able to detect the flux density predicted by a rate of 10$^{-5}\\times$ or less. Ulvestad et al. (2007) estimate the fraction of the Bondi rate of just under 1\\% for G1, but it is difficult to interpret due to the unknown radiative efficiency. For galactic black holes, the radiative efficiencies appear to vary greatly with some lower than $10^{-5}$ (Lowenstein et al. 2001), although consistent with rates of around 10\\% of the Bondi rate. Models which predict 8.6 GHz flux densities from central BHs in globular clusters above about 25 $\\mu \\rm Jy$/beam can be tested with the VLA currently. The EVLA should produce, for continuum observations, a sensitivity improvement of about a factor of 15, making 8.6 GHz flux densities above about 2$\\mu\\rm Jy$/beam detectable." }, "0710/0710.1804_arXiv.txt": { "abstract": "Low-mass X-ray binaries, recycled pulsars, cataclysmic variables and magnetically active binaries are observed as X-ray sources in globular clusters. We discuss the classification of these systems, and find that some presumed active binaries are brighter than expected. We discuss a new statistical method to determine from observations how the formation of X-ray sources depends on the number of stellar encounters and/or on the cluster mass. We show that cluster mass is not a proxy for the encounter number, and that optical identifications are essential in proving the presence of primordial binaries among the low-luminosity X-ray sources. ", "introduction": "The first celestial maps in X-rays, in the early 1970s, show that globular clusters harbour more X-ray sources than one would expect from their mass. As a solution to this puzzle it was suggested that these bright ($L_x\\gtap10^{36}$\\,erg/s) X-ray sources, binaries in which a neutron star captures mass from a companion star, are formed in close stellar encounters. A neutron star can be caught by a companion in a tidal capture, or it can take the place of a star in a pre-existing binary in an exchange encounter. Verbunt \\&\\ Hut (1987) showed that the probability of a cluster to harbour a bright X-ray source indeed scales with the number of stellar encounters occurring in it; whereas a scaling with mass does not explain the observations. With the \\textit{Einstein} satellite a dozen less luminous ($L_x\\ltap10^{35}$\\,erg/s) X-ray sources were discovered in the early 1980. \\textit{ROSAT} enlarged this number to some 55, and now thanks to \\textit{Chandra} we know hundreds of dim X-ray sources in globular clusters. The nature and origin of these dim sources is varied. Those containing neutron stars, i.e.\\ the quiescent low-mass X-ray binaries in which a neutron star accretes mass from its companion at a low rate and the recycled or millisecond radio pulsars, have all formed in processes involving close stellar encounters. The magnetically active binaries, on the other hand, are most likely primordial binaries, with stars that are kept in rapid rotation via tidal interaction. Cataclysmic variables are binaries in which a white dwarf accretes matter from a companion. In globular clusters they may arise either via stellar encounters, or from primordial binaries through ordinary binary evolution -- this is expected to depend on the mass and density of the globular cluster. In this paper we describe the classification and identification of the dim sources in Section\\,2, and make some remarks on the theory of their formation in Section\\,3. In Section\\,4 we will discuss a new, and in our view more accurate, way to compare the numbers of these sources with theoretical predictions. \\begin{figure} \\centerline{ \\parbox[b]{0.55\\columnwidth}{\\psfig{figure=verbuntf1a.eps,width=0.54\\columnwidth,clip=t}} \\parbox[b]{0.45\\columnwidth}{\\psfig{figure=verbuntf1b.ps,width=0.44\\columnwidth,clip=t}} } \\caption{Left: X-ray hardness-luminosity diagram for dim sources in globular clusters. I: quiescent low-mass X-ray binaries, II: cataclysmic variables III: cataclysmic variables and magnetically active binaries. From Pooley \\&\\ Hut (2006). Right: Colour-magnitude diagram of NGC\\,6752 on the basis of HST-WFPC2 data; objects within X-ray position error circles are marked. Left of the main sequence we find cataclysmic variables, above it active binaries. Updated from Pooley et al.\\ (2002a).\\label{xcol}} \\end{figure} ", "conclusions": "\\begin{itemize} \\item mass $M$ is {\\em not} a proxy for collision number $\\Gamma$ \\item the number of dim sources scales both with collision number $\\Gamma$ and with mass $M$ \\item scaling with mass only is not acceptable \\item correct treatment of the background is important, esp.\\ for faint sources \\item to prove the mass-dependence optical identifications are essential \\end{itemize}" }, "0710/0710.3030_arXiv.txt": { "abstract": "We present the results of the timing analysis of five {\\it Rossi X-ray Timing Explorer} observations of the Black Hole Candidate GRS 1915+105 between 1996 September and 1997 December. The aim was to investigate the possible presence of a type-B quasi-periodic oscillation (QPO). Since in other systems this QPO is found to appear during spectral transitions from {\\it Hard} to {\\it Soft} states, we analyzed observations characterized by a fast and strong variability, in order to have a large number of transitions. In GRS 1915+105, transitions occur on very short time scales ($\\sim$ sec): to single them out we averaged Power Density Spectra following the regular path covered by the source on a 3D Hardness-Hardness-Intensity Diagram. We identified both the type-C and the type-B quasi-periodic oscillations (QPOs): this is the first detection of a type-B QPO in GRS 1915+105. As the spectral transitions have been associated to the emission and collimation of relativistic radio-jets, their presence in the prototypical galactic jet source strengthens this connection. ", "introduction": "\\label{par:intro} Systematic variations in the energy spectra and intensity of transient Black-Hole Candidates (BHC) have been recently identified in terms of the pattern described in an X-ray Hardness-Intensity diagram (HID, see Homan et al. 2001, Homan et al. 2005b, Belloni et al. 2005). Four main bright states (in addition to the quiescent state) have been found to correspond to different branches/areas of a square-like HID pattern. In this framework much importance is given to the intermediate states (called Hard Intermediate State, HIMS, and Soft Intermediate State, SIMS) and to the transitions between them, identified from the behaviour in several bands of the electromagnetic spectrum (from radio to hard X-rays, see also Fender, Belloni \\& Gallo 2004 and Homan et al. 2005b) and from the timing properties of the X-ray light curve. Low-frequency Quasi-Periodic Oscillations (LFQPOs) with centroid frequency ranging from mHz to tens of Hz have been observed in the X-ray flux of many galactic BHCs since the '80s (see van der Klis 2006; McClintock \\& Remillard 2006 and references therein). Three main types of LFQPOs, dubbed Type-A, -B and -C respectively, originally identified in the light curve of XTE J1550-564 (Wijnands et al. 1999; Remillard et al. 2002), have been seen in several sources (see Casella et al. 2005 and references therein). We summarize their properties in Table \\ref{ABC_properties}. In the context of the state classification outlined above, it is possible to ascribe the three LFQPOs to different spectral conditions (see Table \\ref{ABC_properties}, Homan et al. 2001, Homan \\& Belloni 2005, Belloni et al. 2005). The type-C QPO is associated to the (radio loud) HIMS and to the low/hard state. It is a common QPO seen in almost all BHCs with a variable centroid frequency correlated with the count rate, a high fractional variability and a high coherence ($Q=\\nu/$FWHM$\\sim$10). The type-B QPO has been seen only in few systems, although it is being seen in a growing number of sources (see Casella et al. 2005 and references therein). It is a transient QPO associated to spectral transitions from the (radio loud) HIMS to the (radio quiet) SIMS. Its features are a $\\sim$ fixed centroid frequency (around $\\sim$6 Hz), lower fractional variability and $Q$ than type-C. Some authors (Fender, Belloni \\& Gallo 2004; Casella et al. 2004) suggested that these spectral transitions are in turn associated to the emission and collimation of transient superluminal relativistic jets visible in radio band. These jets are seen in a number of sources (GRS 1915+105, XTE J1550-564, GX 339-4, XTE J1859+226, GRO J1655-40, etc.). However, not in all of these sources we could resolve the spectral transition to see the transient QPO. \\begin{table} \\label{tab:ABC} \\centering \\caption{Summary of type-A, -B and -C LFQPOs properties (from Casella et al. 2005).} \\label{ABC_properties} \\scriptsize \\begin{tabular}{lccc} \\hline \\hline Properties & Type-C & Type-B & Type-A \\\\ \\hline Frequency (Hz) & $\\sim0.1-15$ & $\\sim5-6$ & $\\sim8$ \\\\ Q($\\nu /$FWHM) & $\\sim7-12$ & $\\gtrsim6$ & $\\lesssim3$ \\\\ Amplitude (\\% {\\it rms}) & 3-16 & $\\sim2-4$ & $\\lesssim3$ \\\\ Noise & strong flat-top & weak red & weak red \\\\ Phase lag$^{\\mathrm{a}}$ @$\\nu_{QPO}$ & soft$/$hard$^{\\mathrm{b}}$ & hard & soft \\\\ Phase lag @2$\\nu_{QPO}$ & hard & soft & ... \\\\ Phase lag @$\\nu_{QPO}/2$ & soft & soft & ... \\\\ \\hline \\end{tabular} \\begin{list}{}{} \\item[$^{\\mathrm{a}}$] With ``hard lag'' we mean that hard variability lags the soft one. \\item[$^{\\mathrm{b}}$] Trend towards soft lags for increasing QPO frequencies \\end{list} \\end{table} The spectral properties connected to type-A QPO are similar to those introduced for the type-B. This QPO has been seen in few systems (Casella et al. 2005). It is broader, weaker and less coherent than the type-B QPO. GRS 1915+105 is a transient BHC discovered on August 15 1992 with the WATCH instrument on board GRANAT (Castro-Tirado et al. 1992, 1994). It is the first galactic source observed to have apparently superluminal transient relativistic radio jets (Mirabel \\& Rodriguez 1994), commonly interpreted as ejection of ultra-relativistic plasma, with a speed close to the speed of light (up to $\\sim98$\\%). Its radio variability was discovered to correlate with the hard X-ray flux (Mirabel et al. 1994). Thanks to VLA-radio observations of these jets, Rodriguez et al. (1995) estimated a distance of 12.5 kpc; more recent estimates attest a distance of $6.5 \\pm 1.6$ kpc (Kaiser et al. 2005). The mass of the compact object was estimated through IR spectroscopic studies (Greiner, Cuby \\& McCaughrean 2001, Harlaftis \\& Greiner 2004) to be $14.0 \\pm 4 M_{\\odot}$ which unambiguously makes GRS 1915+105 a BHC. The various and rich phenomenology of this source was classified by Belloni et al. (2000): they analyzed 163 RXTE observations, showing that the complex behaviour of GRS 1915+105 can be described in terms of spectral transitions between three basic states, A, B and C (not to be confused with the name of the LFQPOs introduced before), that give rise to 12 variability classes. The non standard behaviour of GRS 1915+105 (it is a very bright transient source continuing the same outburst started in 1992) was interpreted as that of a source that spends all its time in Intermediate States (both in its hard and soft flavors), never reaching the LS or the quiescence (see e.g. Fender \\& Belloni, 2004). GRS 1915+105 also shows strong time variability on time scales of fractions of second, revealing low- and high-frequency QPOs (see Morgan et al. 1997) whose properties (frequency and fractional variability) are tightly correlated with the spectral parameters (Morgan et al. 1997; Muno et al. 1999; Markwardt et al. 1999; Rodriguez et al. 2002a, 2002b; Vignarca et al. 2003). In particular, all LFQPOs observed from this system can be classified as type-C QPOs. Although GRS 1915+105 makes a large number of fast state transitions, which have been positively associated to radio activity and jet ejections, no type-B QPO has been observed to date. \\begin{table*} \\centering \\caption{Log of the 5 RXTE/PCA observations analyzed in this work} \\label{log_obs} \\begin{tabular}{c c c c c c} \\hline \\hline N$^{\\circ}$ & Obs. Id.& Date & Starting MJD & Exp. (s) & Classification\\\\ \\hline 1 & 10408-01-35-00 & 1996 Sep 09 & 50348.271 & 9448& $\\mu$\\\\ 2 & 20402-01-45-03 (\\#1, 2, 3) & 1997 Sep 09 & 57700.250 & 10038& $\\beta$\\\\ 3 & 20402-01-53-00 & 1997 Oct 31 & 50752.013 & 9656& $\\beta$\\\\ 4 & 20402-01-53-01 \\& 20402-01-53-02(\\#1) & 1997 Nov 04-05 & 50756.412 & 6322&$\\mu$\\\\ 5 & 20402-01-59-00 & 1997 Dec 17 & 50799.091 & 9784& $\\beta$\\\\ \\hline \\end{tabular} \\flushleft Obs. 4 is composed of two orbits from two separate observations\\\\ \\# indicates the number of the RXTE orbit within the observation\\\\ Classification is from Belloni et al. (2000). \\end{table*} In this paper we present the discovery with RXTE of the type-B QPO in the X-ray light curve of GRS 1915+105. The QPO was present during fast spectral transitions that we identify with the HIMS to SIMS transition observed in other BHCs. ", "conclusions": "\\label{par:discussion} We analyzed 5 RXTE/PCA observations collected during the first two years of the mission. In the power density spectra of all five observations we detected several peaks which we identify with two different types of QPO already seen in many other BHCs: the type-C and type-B QPOs. This is the first identification of a type-B QPO in GRS 1915+105. To detect it, we looked in detail at spectral transitions in observations characterized by a fast and intense variability. Spectral transitions in GRS 1915+105 are usually very fast, often occurring on timescales of $\\sim$ seconds. This is at variance with most of other black-hole binaries in which spectral transitions are observed to last hours or days. In order to study the spectral transitions in GRS 1915+105 we performed an energy-dependent timing analysis by averaging power spectra on the pattern the source recursively tracks in the 3-dimensional hardness-hardness-intensity diagram. Applying this method, we found a type-B QPO in all five observations. In all of them, we detected the type-B QPO together with the type-C, in region \\# 3 of the HHID (see Tab. \\ref{tab:zone}). In three of them, we also detected a type-B QPO alone, in region \\# 4 of the HHID. Two of these observations belong to class $\\mu$ and one to class $\\beta$, which excludes any relation between the class of variability and the presence of the type-B QPO. This means that the presence of the long hard intervals (which differentiate the $\\beta$-class from the $\\mu$-class light curves) does not influence the fast timing properties of the source outside these intervals. We could not find any property (as e.g. hardness, rate) correlated with the presence of the type-B QPO alone in region {\\em 4} of the HHID. In three observations we also found a type-B bump in region {\\em 2}. No correlations were found neither between the presence of the type-B bump and the type-B QPO in region {\\em 4} nor with the hardness and the count rate. \\begin{figure} \\begin{tabular}{c} \\resizebox{7cm}{!}{\\includegraphics{LAG_ALL_DEFINITIVO.ps}} \\end{tabular} \\vspace{-2.0cm} \\caption{Phase lags of the detected QPOs in all observations. For each QPO we extracted the phase lag in a range centered at the QPO peak frequency and corresponding to the width itself ($\\nu_p \\, \\pm \\, FWHM/2$). Errors bar on the X-axis are not shown for clarity.} \\label{fig:lag_all} \\end{figure} \\subsection{QPOs identification} \\label{subpar:disc_identif} In order to identify the two types of QPO that we found in our data sets, we compare them with the known LFQPOs in BHCs. In particular we first analyze their position and behaviour in the Hardness-Intensity diagram (right panel of Figure \\ref{fig:cd_hid_arrows}). When the source moves through regions \\#{\\em 1} and \\#{\\em 2} up to region \\#{\\em 3} the first QPO shows a behaviour very similar to that of type-C QPOs: its frequency is correlated with the count rate and is inversely correlated with the hardness. At a certain hardness, this QPO disappears. A second type of QPO is also detected: as in the case of type-B QPOs, this second QPO appears in a narrow frequency range (often around $\\sim$6 Hz) and in a limited range in hardness. Its frequency and quality factor appear to be slightly correlated with the count rate, particularly when at its lowest frequencies (2.44-2.84 Hz). We interpret this as the first evidence of an increase of the coherence of the type-B QPO from $Q<2$ when at low frequencies to $Q>2$ when reaching frequencies around $\\sim$6 Hz, possibly suggesting the presence of a resonance at this frequency (see Casella et al. 2004). The combined evolution of the QPOs in GRS 1915+105 is strongly reminiscent of the known behaviour of type-C and type-B QPOs in BHCs (see Casella et al. 2005 and references therein). To verify this identification, we plot in Figure \\ref{fig:rms_1} the {\\it rms} fractional variability of the detected QPOs as a function of their frequency for three energy bands. The two QPO types have a somewhat similar energy dependency (being stronger at high energy) but they clearly show different behaviours in these diagrams. In each of the three panels, two well-identified groups of points are evident. A comparison of these two groups with Figure \\#3 of Casella et al. 2004 helps to classify the observed QPO in GRS 1915+105: the first group is diagonally spread across the plots, covering the whole frequency range between $\\sim2$ and $\\sim15$ Hz and a large range in {\\it rms} (particularly at high energies, see the right panel of Figure \\ref{fig:rms_1}). The second group is clustered both in frequency (between $\\sim2.5$ and $\\sim7$ Hz) and in fractional {\\it rms}. The observed behaviour is clearly consistent with that known to be typical of type-C and type-B QPOs in BHCs (see Casella et al. 2004, 2005). \\subsection{Phase lags} \\label{subpar:disc_lag} The association between QPO-type and phase lag in literature is not conclusive: although an average behaviour can be identified (see Tab. \\ref{tab:ABC}, Casella et al. 2005 and reference therein) there are a number of exceptions (see e.g. Belloni et al. 2005 and Homan et al. 2005a). Nevertheless we performed a phase-lag analysis in order to have a comprehensive view of the behaviour of the QPOs in GRS 1915+105. On the basis of the analyzed data it is not possible to characterize unambiguously the phase lag behaviour of any of the two types of QPO we observe: lags appear to vary between different observations, although both types show in average negative values of phase lag (see Fig. \\ref{fig:lag_all}). This can possibly due to the fast variability of the analyzed light curves: we extracted power spectra over time intervals 2 seconds long, without applying any procedure to detrend the variability on longer time scales. However, this variability is very strong (see Fig. \\ref{fig:mu_beta_class}), which results in a leakage at higher frequencies as strong as to actually dominate the phase-lag continuum. \\subsection{GRS 1915+105 as a ``normal'' source} \\label{subpar:disc_normal} The type-B QPO was detected in a few BHCs (see Casella et al. 2005), and associated to spectral transitions from HIMS to SIMS (Homan \\& Belloni 2005, Belloni et al. 2005 and reference therein). Some authors (Gallo et al 2003; Fender, Belloni \\& Gallo 2004) suggested a relation between these spectral transitions and the emission and collimation of transient relativistic radio jets: if we consider the type-B QPO as the signature of radio jet emission, we have this ``signature'' also in the prototypical galactic jet source. As already pointed out by Belloni et al. (2005), the non-detection of a type-B QPO in GRS 1915+105 was rather interesting, especially if you interpret the X-ray$/$radio correlation in this source and in other transients in the framework of the same model (Fender, Belloni \\& Gallo 2004). In GRS 1915+105 the association between X-ray and radio activity is well known (Pooley \\& Fender 1997, Mirabel et al. 1998, Fender \\& Belloni 2004 and references therein). Belloni et al. 2005 suggested that the elusiveness of this QPO (in GRS 1915+105) could be due to high velocity of the movement of the source through the HID. The result presented in this work thus strengthens the interpretation of GRS 1915+105 in the framework of the same model of other BHCs: a type-B QPO appears in correspondence of spectral transition from the B-state to the A-state (Transition, Region {\\em 3}, Stars in Figure \\ref{licu_symbols}) and when the source is in the A-state (Region {\\em 4}, Upward Triangles in Figure \\ref{licu_symbols}). In the light curve in Figure \\ref{licu_symbols} we see state oscillations CTAB TAB TAB that we can identify as fast passages between the HIMS and the SIMS of the other BHCs (Casella et al. 2004; Fender, Belloni \\& Gallo 2004). The type-B QPO appears also in correspondence of the transition from the B-state to the C-state (Squares-Stars-Circles in Figure \\ref{licu_symbols}), therefore not in an oscillation event but in a transition from a soft to a hard state. Furthermore, GRS 1915+105 is at present the heaviest black hole (although uncertainties on BH mass values are rather large, see McClintock \\& Remillard 2006 and references therein) in which a type-B QPO has been observed at 6 Hz (when its quality factor $Q>2$). This strengthens the already known independence of this type of QPO on the mass of the compact object (see Casella et al. 2005). \\subsection{X-ray$/$radio association} \\label{subpar:disc_radio} Klein-Wolt et al. (2002) made a detailed study of simultaneous radio$/$X-ray observations of GRS 1915+105, focusing in particular on radio oscillation events, and found that $\\beta$ class observations are usually associated to strong radio oscillations, while $\\mu$ class observations are usually associated to a weak steady radio activity. These authors point out that the long hard intervals (which are the main macroscopic difference between the two variability classes) appear to be directly linked to the production of radio oscillations. We found the type-B QPO in both classes $\\beta$ and $\\mu$. We could not find any timing property (on time scales shorter than 2 seconds) clearly differentiating the two classes. This apparently weakens the link between the presence of the type-B QPO and radio activity. Unfortunately, the lack of radio coverage during the analyzed RXTE observations does not allow any conclusive result. In order to clarify this issue it will be fundamental to extend our analysis to RXTE observations for which a radio coverage is available." }, "0710/0710.3340_arXiv.txt": { "abstract": "We report the first determination of a distance bracket for the high-velocity cloud (HVC) complex~C. Combined with previous measurements showing that this cloud has a metallicity of 0.15 times solar, these results provide ample evidence that complex~C traces the continuing accretion of intergalactic gas falling onto the Milky Way. Accounting for both neutral and ionized hydrogen as well as He, the distance bracket implies a mass of 3--14\\tdex6~\\Msun, and the complex represents a mass inflow of 0.1--0.25~\\Msunpyr. We base our distance bracket on the detection of \\CaII\\ absorption in the spectrum of the blue horizontal branch star SDSS\\,J120404.78+623345.6, in combination with a significant non-detection toward the BHB star BS\\,16034-0114. These results set a strong distance bracket of 3.7--11.2~kpc on the distance to complex~C. A more weakly supported lower limit of 6.7~kpc may be derived from the spectrum of the BHB star BS\\,16079-0017. ", "introduction": "\\par The evolution of galaxies is strongly driven by the gas in the interstellar medium. There is strong evidence for the infall of new material that provides fuel for galaxy growth. This gas may originate in accreted satellite galaxies, as gas tidally pulled out of passing galaxies, or from pristine intergalactic gas. The cool, infalling clouds appear to be embedded in an extended (100--200~kpc radius) hot Corona (Sembach et al.\\ 2003). Indirect evidence for infalling gas is provided by two arguments: (a) At the current rate of star formation, all of the ISM will be turned into stars within about a Gyr. (b) The narrowness of the distribution of metallicities of long-lived stars implies that the metallicity of the ISM remains more or less constant over a Hubble time, which can happen if there is a continuing inflow of low-metallicity material with a present-day rate of about 1~\\Msunpyr. Item (b) is known as the ``G-dwarf problem'' (van den Bergh 1961). Using the infall hypothesis to solve it has been the subject of much theoretical work (see e.g.\\ Pagel 1997 for a good summary). Continuing infall is essential in detailed numerical modeling of the chemical evolution of the Galaxy and the development of abundance gradients (e.g.\\ Chiappini et al.\\ 2001 and references therein). Infall of low-metallicity gas also seems necessary to reproduce the relatively high abundance of deuterium measured in the local interstellar medium (Linksy et al.\\ 2006). \\par Direct observational evidence for infalling low-metallicity gas is provided by the high-velocity clouds (HVCs; see reviews by Wakker \\& van Woerden 1997; Richter 2006). Subsolar metallicities have now been determined for eleven clouds (see van Woerden \\& Wakker 2004 for a summary). In particular, the metallicity of complex~C is well established as 0.15 times solar (see summary by Fox et al.\\ 2004). Complex~C also has a high deuterium abundance (Sembach et al.\\ 2004). Distance brackets have been more elusive, with just one known before 2006 (8--10~kpc for complex~A -- van Woerden et al.\\ 1999a; Wakker et al.\\ 2003). Thom et al.\\ (2006) derive an 8.8~kpc upper limit for cloud WW\\,35, while in a separate paper (paper~I, Wakker et al.\\ 2007), we present new results for two HVCs (9.8--15.1~kpc for complex~GCP and 5.0--11.7~kpc for the Cohen Stream). In this letter we report a distance bracket for the HVC covering the largest sky area -- complex~C. We summarize our method in Sect.~2. The data are described in Sect.~3, the results in Sect.~4, while in Sect.~5 we summarize the implications. ", "conclusions": "\\par Forty years after the first attempt (Prata \\& Wallerstein 1967), we report the first successful detection of interstellar \\CaII\\ H and K absorption from HVC complex~C. This sets an upper limit on the distance of core CIII (left side in Fig.~1) of 11.2~kpc. For core CI (right side in Fig.~1) we find a lower limit of 3.7~kpc, possibly 6.7~kpc. Although the stars are 27\\deg\\ apart on the sky, it is still safe to conclude that complex~C is located at Galactocentric radius $<$14~kpc, and lies high above the Galactic plane ($z$=3--9~kpc). A more precise determination requires a lower limit for core CIII and an upper limit for CI. \\par Integrating $N$(\\HI) across the cloud, we estimate $M$(\\HI) as 0.7--6\\tdex{6}~\\Msun. \\Ha\\ emission has also been detected (Tufte et al.\\ 1998). We can assume either that the H$^+$ and \\HI\\ are thorougly mixed or that the H$^+$ originates in a photoionized skin around the cloud. In either case, the observed \\Ha\\ intensity suggests that there is roughly as much ionized as neutral gas. \\par We can also estimate the mass inflow associated with complex~C, using a method described in paper~I. Including the neutral and ionized hydrogen, as well as a 40\\% contribution from helium, we derive that complex~C represents about 0.1--0.25~\\Msunpyr\\ of infalling gas. This is a substantial fraction of the theoretically required amount of 1~\\Msunpyr. Other HVCs may contribute the rest, but we have not yet determined distances and metallicities for the most likely candidates. \\par From our results, we conclude that the mystery of the distances to the HVCs is beginning to be solved. The evidence shows that several HVCs are located in the upper reaches of the gaseous Galactic Halo and that they contribute significantly to the inflow of metal-poor gas onto the Galaxy. Once the mass inflow rate is constrained from observations of a sufficient number of HVCs, the next step will be to determine their three-dimensional structure, so that we can use their velocities and galactic location to derive orbits and solve the outstanding mystery of their ultimate origins. \\bigskip Acknowledgements \\par B.P.W., D.G.Y, R.W. and T.C.B. acknowledge support from grant AST-06-07154 awarded by the US National Science Foundation. T.C.B. also acknowledges NSF grants AST 04-06784 and PHY-02-16783; Physics Frontier Center/Joint Institute for Nuclear Astrophysics (JINA). \\par Some of the data presented were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation. \\par The William Herschel telescope is operated on the Island of La Palma by the Isaac Newton group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrophysica de Canarias. \\par Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington." }, "0710/0710.1345_arXiv.txt": { "abstract": "The results obtained from a study of the mass distribution of 36 spiral galaxies are presented. The galaxies were observed using Fabry-Perot interferometry as part of the GHASP survey. The main aim of obtaining high resolution H$\\alpha$ 2D velocity fields is to define more accurately the rising part of the rotation curves which should allow to better constrain the parameters of the mass distribution. The H$\\alpha$ velocities were combined with low resolution HI data from the literature, when available. Combining the kinematical data with photometric data, mass models were derived from these rotation curves using two different functional forms for the halo: an isothermal sphere and an NFW profile. For the galaxies already modeled by other authors, the results tend to agree. Our results point at the existence of a constant density core in the center of the dark matter halos rather than a cuspy core, whatever the type of the galaxy from Sab to Im. This extends to all types the result already obtained by other authors studying dwarf and LSB galaxies but would necessitate a larger sample of galaxies to conclude more strongly. Whatever model is used (ISO or NFW), small core radius halos have higher central densities, again for all morphological types. We confirm different halo scaling laws, such as the correlations between the core radius and the central density of the halo with the absolute magnitude of a galaxy: low luminosity galaxies have small core radius and high central density. We find that the product of the central density with the core radius of the dark matter halo is nearly constant, whatever the model and whatever the absolute magnitude of the galaxy. This suggests that the halo surface density is independent from the galaxy type. ", "introduction": "Rotation curves are a fundamental tool for studying the dynamics and mass distribution in galaxies. The distribution of the total mass can then be compared with the distribution of visible light, assuming a certain mass-to-light (M/L) ratio. Observations have clearly established that dark matter is needed for explaining the rotation velocities observed in the outer parts of spirals. The dark matter is most often considered as being distributed in a spherical dark halo but its density profile, especially at small radii, is still a matter of debate (Blais-Ouellette et al. 2001, de Blok \\& Bosma 2002, Swaters et al. 2003, Navarro 2004, Graham et al. 2006, Kusio de Naray 2006, Hayashi et al. 2007). However, mass models of spiral and dwarf galaxies have well-known degeneracies (Barnes et al. 2004) that prevent a unique mass decomposition, the most important being the unknown value of the stellar mass-to-light ratio (Dutton et al. 2005). Unfortunately, stellar population models (e.g. de Jong \\& Bell 2006) still cannot predict accurately the (M/L) values of the stellar disks based on their colors. The main problem comes from the fact that the normalization of this relation (color vs M/L) depends critically on the shape of the stellar IMF at the low mass end. It is well known that the faint stars contribute significantly to the mass but not to the luminosity and color of the stellar disks. This means that it will prove difficult to tighten this relation and lessen the effect of the disk-halo degeneracy. Cosmological numerical simulations favor cuspy dark halos, although the value of the inner slope $\\gamma$ of the radial density profile, where $\\rho$ $\\propto$ r $^{\\gamma}$ (see equation 6 of section 4) depends on the authors, with $\\gamma$ = -1.5 for Moore et al. 1999 as well as Fukushige \\& Makino 2001 or -1.0 for Navarro, Frenk \\& White 1997 (hereafter NFW). Recent simulations (Graham et al. 2006) have extrapolated inner logarithmic profile slope ranging from -0.2 to -1.5, with a typical value at 0.1 kpc around -0.7. Most observers conclude however that $\\gamma$ is closer to 0.0 than to -1.0 and that mass models give better results with an isothermal (or pseudo-isothermal) sphere halo rather than a NFW profile (e.g. Blais-Ouellette et al. 2001, de Blok \\& Bosma 2002, Swaters et al. 2003, de Blok et al. 2003, de Blok 2005, Kassin et al. 2006a). Most of these observations are based on the rotation curves of dwarf and Low Surface Brightness (LSB) galaxies (Kuzio de Naray et al. 2006). It is still unclear if this is true also for High Surface Brightness (HSB) systems and for all morphological types. One problem in this cusp-core debate is that numerical simulations mainly predict the halo density profile shape at the time of formation while galaxies are observed after many Gyrs of evolution. The problem is that internal dynamics and interaction between the dark halo and the luminous disk (e.g. adiabatic contraction: Dutton et al. 2005) and interaction with the environment (Maccio et al. 2007) may have altered the shape of the halo density profile. Moreover, the shape of the gravitational potential in CDM halos may explain the core-like halo density profiles seen in LSB systems. Hayashi et al. 2007 suggest that galactic disks may be forming in elliptical gravitational potential. This could result in significant non-circular motions in systems such as LSBs which would mimic constant density cores. Thus, taking into account the 3D shape of the dark mass distribution could reconcile the constant density cores observed in LSBs with the predicted cuspy mass profiles of CDM halos. Finally, it is unclear if CDM simulations, mainly obtained to compare with the large scale structure, have sufficient resolution to reliably probe the kpc scales necessary to compare with the observationally derived dark matter halo density profiles on the scale of a few halo core radii. As discussed by Navarro (2004), unfortunately rotation curves constraints are strongest where numerical simulations are the least reliable. In fact, rotation curves are usually compared with extrapolations of the simulation data that rely heavily on the applicability and accuracy of fitting formulae such as the NFW profile to regions that may be compromised by numerical artifact. This is especially true for LSB and dwarf galaxies (for instance Dutton et al. 2005). Blais-Ouellette et al. 1999 and Barnes et al. 2004 have shown the necessity of optical integral field spectroscopy to accurately determine the rotation curves in the inner parts of spiral and dwarf galaxies, for which the HI data are affected by beam smearing (Swaters et al. 2000, van den Bosch et al. 2000). While the use of two-dimensional data does not necessarily alter the halo parameters derived from optical long-slit data it however decreases the uncertainties by roughly a factor of 2 (de Naray et al. 2006). Blais-Ouellette et al. 2001 and Blais-Ouellette et al. 2004 also pointed out the great sensitivity of the mass distribution parameters to the inner rotation curve. The ideal rotation curve will therefore combine high resolution optical data, for the inner part, with radio data, for the outer part extending well beyond the luminous disk. ", "conclusions": "Our study, based on 36 galaxies of different morphological types, confirms the result already claimed by other authors about the shape of the dark matter halo in the center of spiral galaxies, namely that its density profile is probably closer to an isothermal sphere profile than to an NFW profile: DM halos are rather flat than cuspy. Our mass models (ISO and NFW) are compared when fitting the rotation curve as a whole, however a comparison of the chi2 values limited to the central regions would be still clearer since the inner slope of the RC found with the NFW model is systematically steeper and above the data points, compared with the ISO model. Thus far, most of the results found in the literature concerned dwarf or Low Surface Brightness galaxies. 9 galaxies of our sample are LSB (UGC 2304, 5272, 5279, 5789, 7524, 7699, 9465, 11707 and 12060, with types going from Scd to Im) and, for all of them, we systematically find better results with the ISO model (smaller $\\chi^{2}$) than with NFW, except UGC 2304 for which both are equivalent. Interestingly, our study suggests that this holds for most spirals since the 36 galaxies of this study have morphological types ranging from Sab to Im. For almost all of them, the best fit is obtained with the ISO model (the reverse is found only for 4 galaxies out of 36). Also, no significant difference can be seen when comparing the quality of the fits obtained with the NFW and the ISO model as a function of the morphological type. However, a difference can be seen in the way the rotation curves are decomposed into several components, with the halo being the main component for late types and the disk (or disk + bulge) the main component for early types. This result, although needing to be confirmed with a larger sample, merely reflects the well known fact that later type galaxies are more dark matter dominated. Finally, we confirm different halo scaling laws seen previously by other authors such as C\\^ot\\'e et al (2000), Kormendy \\& Freeman 2004 and Barnes et al. 2004. Among those, it appears clearly that low luminosity galaxies have small core radius and high central density, the product of the two parameters being nearly constant with absolute magnitude. This means that the galaxy halo surface density is independent of galaxy type or luminosity. Trends are also seen for the core radius as a function of luminosity and color but should be confirmed with a larger sample." }, "0710/0710.4560_arXiv.txt": { "abstract": "We study the effect of primordial nongaussianity on large-scale structure, focusing upon the most massive virialized objects. Using analytic arguments and N-body simulations, we calculate the mass function and clustering of dark matter halos across a range of redshifts and levels of nongaussianity. We propose a simple fitting function for the mass function valid across the entire range of our simulations. We find pronounced effects of nongaussianity on the clustering of dark matter halos, leading to strongly scale-dependent bias. This suggests that the large-scale clustering of rare objects may provide a sensitive probe of primordial nongaussianity. We very roughly estimate that upcoming surveys can constrain nongaussianity at the level $|\\fnl|\\lesssim 10$, competitive with forecasted constraints from the microwave background. ", "introduction": "One of the fundamental predictions of standard (single-field, slow-roll) inflationary cosmology is that the density fluctuations in the early universe that seeded large-scale structure formation were nearly gaussian random (e.g.\\ \\cite{maldacena,Acquaviva:2002ud,Creminelli:2003iq,Lyth_Rodriguez,Seery_Lidsey}). Constraining or detecting non-gaussianity (NG) is therefore an important and basic test of the cosmological model. To the extent that it can be measured, gaussianity has so far been confirmed; the tightest existing constraints have been obtained from observations of the cosmic microwave background \\cite{wmap3,Creminelli_wmap}. Recently, several inflationary models have been proposed which predict a potentially observable level of nongaussianity, see \\eg \\cite{ArkaniHamed:2003uz,Bartolo:2003jx,Lyth:2005du,Rigopoulos:2005ae, Allen:2005ye,Chen_DBI,Barnaby:2006cq,Barnaby:2006km,Barnaby:2007yb, Sasaki:2006kq,Chen:2006nt,Chen_Easther_Lim,Battefeld_Easther, Assadullahi:2007uw,Battefeld:2007en,Shandera} and \\cite{Bartolo:2004if} for a review. Improved limits on NG would rule out some of these models; conversely, a robust detection of primordial nongaussianity would dramatically overturn standard inflationary cosmology and provide invaluable information about the nature of physical processes in the early universe. In this regard, there has been a resurgence in studying increasingly more sophisticated methods and algorithms to constrain (or, if we are lucky, detect) nongaussianity \\cite{Babich:2005en,Babich_shape,Creminelli_estimators,Smith_Zaldarriaga,Fergusson_Shellard}. Nongaussianity manifests itself not only in the cosmic microwave background \\cite{Falk_Ran_Sre,Luo_Schramm,Gangui_etal,Wang_Kam}, but also in the late-time evolution of large-scale structure. For example, detailed measurements of higher order correlations like the bispectrum or trispectrum of galaxy clustering could provide a handle on primordial nongaussianity \\cite{Verde:2000vr,Scoccimarro:2003wn,Sefusatti:2007ih}. The abundance of galaxy clusters, the largest virialized objects in the universe, has also long been recognized as a sensitive probe of primordial NG \\cite{Lucchin:1987yv,Robinson:1999se,Benson:2001hc,Matarrese:2000iz,verde01,Scoccimarro:2003wn,Komatsu:2003fd}. Because clusters are rare objects which form from the largest fluctuations on the tails of the density probability distribution, their abundance is keenly sensitive to changes in the shape of the PDF such as those caused by nongaussianity. Large statistical samples of massive clusters have already been compiled from wide-area optical imaging and spectroscopic surveys such as the Sloan Digital Sky Survey \\cite{maxBCG,Koester:2007bg}, the Two-Degree Survey \\cite{Eke:2004ve}, and from the Red Sequence Survey \\cite{Yee:2007if} and from X-ray surveys using the Chandra and XMM-Newton observatories \\cite{Willis:2005ag,Valtchanov:2003it}. Future missions, such as the Dark Energy Survey, Supernova/Acceleration Probe and Large Synoptic Survey Telescope, will detect and study tens of thousands of clusters, revolutionizing our understanding of cluster physics as well as providing important constraints on cosmology \\cite{Haiman_Mohr_Holder,Majumdar_Mohr,Wang_Haiman, Battye_Weller,Lima_Hu_05, Marian_Bernstein,Takada_Bridle}. To exploit the potential of these upcoming surveys as probes of primordial nongaussianity, it is important to calibrate the effects of NG on the abundance and clustering of virialized objects. While no previous work has attempted to quantify the effects of NG on halo clustering, several groups over the past decade have constructed fitting formulae for the halo mass function \\cite{Robinson_Baker,Robinson_Gawiser_Silk,MVJ}. All of this work, however, was analytic and relied on the validity of the Press-Schechter \\cite{press-schechter} formalism, plus various further approximations. The resulting analytic estimates are, in general, rather cumbersome to compute and have questionable accuracy. As discussed below, the Press-Schechter model provides only a qualitative description of halo abundance, and fails to reproduce the halo mass function to within an order of magnitude over the mass and redshift range accessible to current and future cluster surveys. Therefore, analytic models for NG cluster abundance based on the Press-Schechter ansatz may not be sufficiently accurate. Given the high-quality data soon to be available, a much more precise calculation of cluster statistics will be required. Quite recently, two groups have attempted to quantify the mass function of clusters in NG models using N-body simulations \\cite{Kang,Grossi}, reaching contradictory conclusions. In this paper, we use analytic arguments and numerical simulations to estimate the effect of NG on the abundance and clustering of virialized objects. Because N-body simulations can be expensive and there is a wide NG parameter space, we also strive to make our results useful to a cosmologist who is not necessarily equipped with the machinery or patience to run simulations or evaluate difficult analytic expressions. To this end, we provide a simple, physically motivated fitting formula for the halo mass function and halo bias, which we calibrate to our N-body simulations. Our main results are that the mass function and correlation function of massive halos can be significantly modified by primordial nongaussianity. We find a somewhat weaker effect of NG on the mass function than previous analytic estimates. We also show analytically and numerically that NG strongly affects the clustering of rare objects on large scales, implying that measurements of the large-scale power spectrum can place stringent bounds on NG. The plan of the paper is as follows. In Section \\ref{anal} we derive analytic expressions for the abundance and clustering of rare peaks. In Section \\ref{sec:method} we describe our N-body simulations, followed in Section \\ref{sec:mf} by a discussion of our measured halo mass function, and our fitting formula for the mass function. In Section \\ref{sec:pk} we present measurements of halo clustering within our simulations, and in Section \\ref{sec:cosmo} we discuss cosmological implications of our findings. ", "conclusions": "We have quantified the effects of primordial nongaussianity on the abundance and power spectra of massive halos. Our two principal results are as follows. First, we have provided a new fitting formula for the halo mass function. The formula is based on matching halos in Gaussian and non-Gaussian simulations: for $\\fnl>0$ the corresponding halos are more massive than in the Gaussian case, and vice versa. The formula is consistent with the measured mass function from our simulations to within $\\sim10\\%$ over the entire range of masses and redshifts that we consider. Being essentially a convolution of the Gaussian mass function and a Gaussian kernel (Eqs.~(\\ref{eq:mf_conv})-(\\ref{eq:rms_Mf_def})), the formula is also easy to use and does not require estimating the extreme tails of the nongaussian PDF of the density field. Our results also indicate that previous work based on Extended Press-Schechter type formulae overestimated the effects of nongaussianity on the abundance of halos by a factor of $\\sim 2$ over the relevant mass scales. Secondly, we showed both analytically and numerically that nongaussianity (in the $\\fnl$ model) leads to strong scale dependence of the bias of dark matter halos. We find remarkably good agreement between our analytic expression and our numerical results. Measurement of the power spectrum of biased objects therefore provides a new avenue to detect and measure nongaussianity. While cluster counts can constrain NG at a level comparable to existing CMB constraints, $|\\fnl|\\lesssim 100$, we found that future large-scale redshift surveys can potentially do much better, roughly $|\\fnl|\\lesssim 10$. We do not find significant degeneracies between $\\fnl$ and dark energy parameters in our Fisher matrix calculations, either for mass function measurements or power spectrum measurements. More precise estimates will require considerably more sophisticated treatments than we have attempted in our illustrative examples above. We close this paper by considering, in light of our findings, the optimal methods for constraining NG of the $\\fnl$ form. Measurements of the power spectrum would appear the most promising; observations of high redshift, highly clustered objects on large scales would allow the strongest constraints on the scale-dependent bias signature of $\\fnl$. Fortunately, upcoming BAO surveys will likely provide the necessary observations of, e.g. luminous red galaxies (LRGs). Photometric surveys may also be useful in this regard. Since the effects of NG are most pronounced on large scales, rather than small scales, precise spectroscopic redshifts may not be necessary. Photometric redshifts with errors of order $\\Delta z\\approx 0.03$ have already been achieved for LRGs and for optically selected groups and clusters with prominent red sequences \\cite{Padmanabhan:2004ic,Ilbert:2006dp,Yee:2007if}. At $z=0.5$, this corresponds to roughly 100 $h^{-1}$Mpc comoving, fairly small compared to the $\\sim$Gpc scales where NG becomes most important. Since photometric surveys can cover wider areas more deeply than spectroscopic surveys, they may turn out to provide tighter bounds. Besides their abundance and clustering, the internal properties of massive halos may also be sensitive to nongaussianity. For instance, the concentrations and substructure content of massive halos have been found to depend upon primordial NG \\cite{avila03}. Our simulations lacked sufficient force resolution to explore this in detail, but we note in passing that multiple groups find a tension between observations of massive lensing clusters and theoretical predictions for Gaussian perturbations \\cite{rcs,arcs04,hennawi07,broadhurst08}. Another intriguing possibility for probing primordial NG is to use statistics of the largest voids in the universe. Just as the abundance and clustering of high density peaks are affected by nongaussianity, so are the same properties for deep voids (albeit with an opposite sign, c.f.\\ Fig.~\\ref{fig:slice_sims}). In a sense, because voids are not as nonlinear as overdense regions, their properties are more easily related to the initial Lagrangian underdensities whose statistics are straightforward to compute. Voids may be detected at high redshift as a deficit of Lyman-$\\alpha$ forest absorption features in QSO spectra. The Sloan Digital Sky Survey (SDSS) has already measured spectra for high redshift QSO's over a roughly $\\sim$8000 deg$^2$ area, corresponding to a volume of $\\gtrsim 30 ({\\rm Gpc}/h)^3$ \\cite{sdss_lyaf}. Each QSO spectrum typically probes $\\sim 400 h^{-1}$ Mpc, and the typical transverse separation between QSO sightlines in SDSS is $\\sim 100 h^{-1}$ Mpc, (P.\\ McDonald, priv.\\ comm.) so measurements of the clustering of $\\sim 10$ Mpc-sized voids on $\\sim$ Gpc scales may already be feasible. Finally, we note that our conclusions are based on simulations implementing a very specific type of local primordial nongaussianity quantified by the $\\fnl$ parameter. The validity of our conclusions in the context of other type of primordial nongaussianity is the subject of ongoing studies." }, "0710/0710.2493_arXiv.txt": { "abstract": "{The annihilations of WIMPs produce high energy gamma-rays in the final state. These high energy gamma-rays may be detected by IACTs such as the H.E.S.S. array of Imaging Atmospheric Cherenkov telescopes. Besides the popular targets such as the Galactic Center or galaxy clusters such as VIRGO, dwarf spheroidal galaxies are privileged targets for Dark Matter annihilation signal searches. H.E.S.S. observations on the Sagittarius dwarf galaxy are presented. The modelling of the Dark Matter halo profile of Sagittarius dwarf is discussed. Constraints on the velocity-weighted cross section of Dark Matter particles are derived in the framework of Supersymmetric and Kaluza-Klein models. The future of H.E.S.S. will be briefly discussed. \\PACS{ {98.70.Rz}{Gamma-rays : observations} \\and {95.35.+d}{Dark Matter} } % } % ", "introduction": "\\label{intro} H.E.S.S. (High Energy Stereoscopic System) is an array of four Imaging Atmospheric Cherenkov Telescopes (IACTs) located in the Khomas Highlands of Namibia at an altitude of 1800 m above sea level. Each telescope has an optical reflector of 107 m$^2$ composed of 382 round mirrors \\cite{bernlohr}. The Cherenkov light is emitted by charged particles from electromagnetic showers initiated by the interaction of the primary gamma-rays in the Earth's upper atmosphere. The light is collected by the reflector which focuses it on a camera made of 960 fast photomultiplier tubes (PMTs) with individual field of view of 0.16$^{\\circ}$ in diameter \\cite{vincent}. Each PMT is equipped with a Winston cone to maximize the light collection and limit the background light. The total field of view of the H.E.S.S. instrument is 5$^{\\circ}$ in diameter. The stereoscopic technique allows for an accurate reconstruction of the direction and energy of the primary gamma-rays as well as for an efficient rejection of the background induced by cosmic ray interactions. The energy threshold is about 100 GeV at zenith and the angular resolution is better than 0.1$^{\\circ}$ per gamma-ray. The point source sensitivity is $\\rm 2 \\times 10^{-13} cm^{-2}s^{-1}$ above 1 TeV for a 5$\\sigma$ detection in 25 hours \\cite{crabe}.\\\\ The H.E.S.S. instrument is designed to detect very high energy (VHE) gamma-rays in the 100 GeV - 100 TeV energy regime and investigate their possible origin. Scenarios beyond the Standard Model of particle physics predict plausible WIMP (weakly interacting massive particle) candidates to account for the Cold Dark Matter. Among these are the minimal supersymmetric extension of the Standard Model (MSSM) or universal extra dimension (UED) theories. With R-parity conservation, SUSY models predict the lightest SUSY particle (LSP) to be stable. In various SUSY breaking scenarios, the LSP is the lightest neutralino $\\rm \\tilde{\\chi}$. In Kaluza-Klein (KK) models with KK-parity conservation, the lightest KK particle (LKP) is stable \\cite{KK}, the most promising being the first KK mode of the hypercharge gauge boson, $\\rm \\tilde{B}^{(1)}$. The annihilation of WIMPs in galactic halos may produce gamma-ray signals detectable with Cherenkov telescopes (for a review, see \\cite{bertone}). Generally, their annihilations will lead to a continuum of gamma-rays with energy up to the WIMP mass resulting from the hadronization and decay of the cascading products, mainly from $\\pi^0$'s generated in the quark jets. The gamma-ray flux from DM particle annihilations of mass m$_{DM}$ in a spherical halo can be expressed as: \\begin{equation} \\frac{d\\Phi(\\Delta\\Omega,E_{\\gamma})}{dE_{\\gamma}} = \\frac{1}{4\\pi}\\,\\frac{\\langle \\sigma v \\rangle}{m^2_{DM}}\\frac{dN_{\\gamma}}{dE_{\\gamma}}\\,\\times\\,\\bar{J}\\Delta\\Omega \\end{equation} which is a product of a particle physics term containing the velocity-weighted cross section $\\langle \\sigma v \\rangle$ and the differential gamma-ray spectrum $dN_{\\gamma}/dE_{\\gamma}$, and an astrophysics term $\\bar{J}$ given by: \\begin{equation} \\bar{J} = \\frac{1}{\\Delta\\Omega}\\int_{\\Delta\\Omega}\\int_{l.o.s}\\rho^2 ds \\end{equation} This term corresponds to the line-of-sight-integrated squared density of the DM distribution which is averaged over the solid angle integration region spanned by the H.E.S.S. point spread function (PSF).\\\\ Plausible astrophysical targets have been proposed to search for DM, from local galactic objects to extragalactic objects such as galaxy clusters. This paper reports on three recent results on targets relevant to indirect dark matter search: the Galactic Center, the center of the VIRGO cluster (M87) and the Sagittarius dwarf spheroidal galaxy. ", "conclusions": "\\label{conclusion} The H.E.S.S. collaboration has studied potential targets for DM annihilations. The TeV $\\gamma$-ray energy spectrum measured by H.E.S.S. in the Galactic Center region is unlikely to be interpreted in terms of WIMP annihilations. Constraints on the velocity-weighted annihilation cross-section have been derived in the case of a NFW DM halo profile. Neither pMSSM nor KK models can be ruled out. The detection of a temporal variability of the VHE signal from the M87 nucleus in the VIRGO galaxy cluster excludes the bulk of the gamma-ray signal to come from DM annihilations. H.E.S.S. observed the Sagittarius dwarf spheroidal galaxy and no significant gamma-ray excess has been detected. Two DM halo modellings of Sgr have been investigated to derive contraints on the velocity-weighted annihilation cross-section. Some models can be ruled out in the case of a cored profile.\\\\ Dark matter searches will continue and searches with H.E.S.S. 2 will start in 2009. The phase 2 will consist of a new large 28 m diameter telescope located at the center of the existing array. With the availability of the large central telescope H.E.S.S. 2, the analysis energy threshold will be lowered down to less than 80 GeV and will allow to explore more supersymmetric models." }, "0710/0710.2941_arXiv.txt": { "abstract": "Although the occurrence of steady supercritical disk accretion onto a black hole has been speculated about since the 1970s, it has not been accurately verified so far. For the first time, we previously demonstrated it through two-dimensional, long-term radiation-hydrodynamic simulations. To clarify why this accretion is possible, we quantitatively investigate the dynamics of a simulated supercritical accretion flow with a mass accretion rate of $\\sim 10^2 L_{\\rm E}/c^2$ (with $L_{\\rm E}$ and $c$ being, respectively, the Eddington luminosity and the speed of light). We confirm two important mechanisms underlying supercritical disk accretion flow, as previously claimed, one of which is the radiation anisotropy arising from the anisotropic density distribution of very optically thick material. We qualitatively show that despite a very large radiation energy density, $E_0\\gsim 10^2L_{\\rm E}/4\\pi r^2 c$ (with $r$ being the distance from the black hole), the radiative flux $F_0\\sim cE_0/\\tau$ could be small due to a large optical depth, typically $\\tau\\sim 10^3$, in the disk. Another mechanism is photon trapping, quantified by ${\\bm v}E_0$, where ${\\bm v}$ is the flow velocity. With a large $|{\\bm v}|$ and $E_0$, this term significantly reduces the radiative flux and even makes it negative (inward) at $r < 70 r_S$, where $r_S$ is the Schwarzschild radius. Due to the combination of these effects, the radiative force in the direction along the disk plane is largely attenuated so that the gravitational force barely exceeds the sum of the radiative force and the centrifugal force. As a result, matter can slowly fall onto the central black hole mainly along the disk plane with velocity much less than the free-fall velocity, even though the disk luminosity exceeds the Eddington luminosity. Along the disk rotation axis, in contrast, the strong radiative force drives strong gas outflows. ", "introduction": "Recently, very bright objects which may be undergoing supercritical (or super-Eddington) accretion flows have successively been found. Good examples are ultraluminous X-ray sources (ULXs; \\citeauthor{WMM01} \\citeyear{WMM01}; \\citeauthor{Ebisawa03} \\citeyear{Ebisawa03}; \\citeauthor{Okajima06} \\citeyear{Okajima06}). These are pointlike off-center X-ray sources whose X-ray luminosity significantly exceeds the Eddington luminosity of a neutron star \\citep{Fabbiano89}. Because of substantial variations, it is reasonable to assume that the ULXs are single compact objects powered by accretion flows \\citep{Makishima00}. If so, there are two possibilities to account for large luminosities exceeding the Eddington luminosity for a mass of 100 $M_\\odot$: sub-critical accretion onto an intermediate-mass black hole (IMBH) and supercritical accretion onto a stellar-mass black hole. We support the latter possibility, since through the fitting to several {\\it XMM-Newton} EPIC data of ULXs, which have been claimed as good IMBH candidates, we have found evidence of supercritical flows \\citep{Vierdayanti06}. Another interesting group is narrow-line Seyfert 1 galaxies \\citep[see][for a review]{Boller04}. Because of their relatively small black hole masses, they have in general large Eddington ratios ($L/L_{\\rm E}$ with $L$ and $L_{\\rm E}$ being the luminosity and the Eddington luminosity, respectively), and some of them seem to fall in the slim-disk regimes (\\citeauthor{Mineshige00} \\citeyear{Mineshige00}; \\citeauthor{Kawaguchi03} \\citeyear{Kawaguchi03}; see also \\citeauthor{Wang99} \\citeyear{Wang99}) Despite growing evidence indicating the existence of supercritical accretion flows in the universe, theoretical understanding is far from being complete. It is well known that any spherically accreting object, irrespective of the nature of the central source, cannot emit above the Eddington luminosity, since otherwise significant radiative force will prevent accretion of the gas. If we examine detailed radiation-matter interactions in the interior, however, we notice that the situation is not so simple. Actually, radiation produced at the very center cannot immediately reach the surface (i.e., photosphere), since photons generated at the center should suffer numerous scatterings with accreting material and thus take a long time to reach the surface. If the matter continuously falls, and if the mass accretion timescale is shorter than the mean travel time for photons to reach the surface (the diffusion timescale), photons at the core may not be able to go out. This is the so-called photon-trapping effect \\citep[e.g.,][]{Begelman78,HC91}. Here we define the trapping radius, inside which radiation-matter interaction is so frequent that photons are trapped within the accretion flow. Inside this trapping radius, therefore, radiative flux can be negative (i.e., inward). Thus, the apparent luminosity is reduced as compared with the case without photon trapping. Nevertheless, the concept of the Eddington luminosity is still valid, since far outside the trapping radius radiative flux should be outward and its absolute value should be less than $L_{\\rm E}/4\\pi r^2$, where $r$ is the distance from the central black hole, in the quasi-steady state. Radiation-hydrodynamic (RHD) simulations of spherically symmetric supercritical accretion flows have been performed by \\citet{BK83}. The situation may differ in the case of disk accretion, since the radiation field is not isotropic due to inhomogeneous matter distribution. That is, matter can fall predominantly along the disk plane, whereas radiation can go out along its rotating axis, where matter density is low. In other words, the main directions of the inward matter flow and the outward radiative flux are not parallel to each other in disk accretion, leading to a situation in which the radiative force does not completely counteract the gravitational force. There is room for the possibility of a supercritical accretion flow with super-Eddington luminosity. Based on such an argument, many researchers have speculated about the occurrence of supercritical flows in disk accretion systems. \\citet{SS73} discussed the possibility of supercritical disk accretion based on a one-dimensional steady model \\citep[see also][]{MRT76}. They mentioned that the mass accretion rate would not be steady but oscillate if the mass accretion rate exceeded the critical rate; otherwise, a part of the accreting matter might be ejected from the disk as a disk wind. Radiatively driven outflows from supercritical disks were investigated by \\citet{Meier79}, \\citet{JR79}, and \\citet{Icke80}. Despite a long history in the study of supercritical disk accretion flows, the occurrence of steady supercritical disk accretion has not yet been accurately verified. Similar simulations have been performed since the 1980s, but all of them calculated only the initial transient phase. Their conclusions then were not general, since the back-reactions (i.e., enhanced radiation pressure), which may inhibit steady flow, were not accurately evaluated. Although the radiative force predominantly has an effect in the vertical direction, it should also have an effect on the material within the disk. Hence, in the direction parallel to the disk plane, the situation may be similar to or more severe than the case of spherical accretion. This is because the radiative force, together with the centrifugal force, may possibly overcome the gravitational force. To study the possibility of supercritical disk accretion flows, precise and quantitative research treating both supercritical disks and outflows, and which also takes into consideration multi-dimensional effects, is needed. \\citeauthor{O05} (\\citeyear[][hereafter Paper I]{O05}) have confirmed the occurrence of quasi-steady supercritical disk accretion onto a black hole by two-dimensional RHD simulations. The motivation of the present study is to investigate the physical mechanisms which make supercritical disk accretion possible. For this purpose we examine quantitatively the flow motion and force fields via the radiative flux, rotation, and gravity of a supercritical disk accretion flow onto a central black hole, based on the two-dimensional RHD simulation data from Paper I. Through detailed inspection of the results, it will be possible to clarify the physics behind supercritical disk accretion flows. In \\S 2 we plot the spatial distributions of several key quantities which control flow dynamics. A discussion is given in \\S 3. ", "conclusions": "\\subsection{Structure of the Supercritical Disk Accretion Flow} We summarize our simulation results in a schematic picture (see Fig. \\ref{pic}). \\begin{figure*} \\epsscale{0.87} \\plotone{Fig07_astroph.eps} \\caption{Schematic picture of the supercritical disk accretion flow around a black hole. The gas motion is shown on the right: the high-velocity outflow ({\\it solid arrows}) and the slow accretion flow ({\\it dashed arrows}). The left side indicates the radiative fluxes in the comoving frame ({\\it black arrows}) and the rest frame ({\\it white arrows}). \\label{pic} } \\end{figure*} In this figure the radiative flux in the comoving frame, $F_0^r$, is positive (outward) except in the vicinity of the black hole. However, the radiative flux in the rest frame, $F^r (\\sim F_0^r+4v_r E_0/3)$, is negative (inward) via photon trapping, $v_r E_0<0$, in the trapping region. In the vicinity of the black hole, both $F_0^r$ and $F^r$ are negative. Matter slowly accretes inside the disk, since the sum of the radiative and centrifugal forces is nearly balanced with the gravitational force. Here we stress again that the radiative force counteracts the gravitational force in spite of the trapping region because $F_0^r>0$. Radiatively driven high-velocity outflows appear above and below the disk. In the very vicinity of the black hole, the gas is accelerated inward by the radiative force and the gravitational force. As far as the physical quantities around the equatorial plane are concerned, the simulated profiles of the density, temperature, and radial as well as rotational velocities roughly agree with the prediction of the slim-disk model \\citep{Abramowicz88}. Such features have already been shown in Figure 11 in Paper I. However, only about $10\\%$ of the injected mass can accrete onto the black hole, and an almost equal amount of matter is ejected as high-velocity outflows. The mass accretion rate is not constant in the radial direction and decreases near the black hole (see Fig. 6 in Paper I). Thus, we conclude that the simulated flows do not perfectly agree with the slim-disk model with regard to the whole structure of the flow. \\citet{HD07} have investigated local maximum values of the accretion rate in the supercritical disk accretion flows. In their paper they focused only on the force balance in the vertical and radial directions around the equatorial plane. Multi-dimensional effects were not taken into consideration. They revealed that the vertical radiative force limits the maximum accretion rate at the inner disk region, leading to a decrease of the accretion rate with a decrease of the radius. Their results imply that the disk loses mass via the radiatively driven outflows and the mass accretion rate decreases at the inner region. Such tendencies agree with our results. As shown in Figure \\ref{vr}, our simulations show that radiatively driven outflows form above and below the disk. The mass accretion rate decreases with a decrease of the radius (see Fig. 6 in Paper I). \\subsection{Dependency of the Mass Accretion Rate} In the present study, focusing on numerical simulations in which the mass input rate at the outer boundary is set to be $\\dot{M}_{\\rm input}=10^3L_{\\rm E}/c^2$, we show that the radiative force is attenuated in the disk region via large optical thickness, which makes supercritical disk accretion possible. Such dilution of the radiative force would effectively operate even if the mass input rate (mass accretion rate) varied. In fact, simulations with mass input rates of $3\\times 10^2L_{\\rm E}/c^2$ and $3\\times 10^3L_{\\rm E}/c^2$ show that $E_0/\\rho$ is almost independent of the mass input rate, although the radiation energy density goes up as the mass input rate increases. That is, the dynamics is not sensitive to the precise value of $\\dot{M}_{\\rm input}$, as long as it largely exceeds the critical value. So far, we have studied steady accretion flows. Although a highly supercritical disk ($\\dot{M}_{\\rm input}>3\\times 10^2L_{\\rm E}/c^2$) is quasi-steady, it has been revealed that a moderately supercritical disk [$\\dot{M}_{\\rm input}=(10-10^2)L_{\\rm E}/c^2$] is unstable and exhibits limit-cycle behavior (\\citeauthor{O06} \\citeyear{O06}, \\citeyear{O07}; see also \\citeauthor{SH75} \\citeyear{SH75}; \\citeauthor{SS76} \\citeyear{SS76}). The luminosity goes up and down around the Eddington luminosity. \\citet{O07} has reported that the time-averaged mass, momentum, and kinetic energy output rates via the outflow, the mass accretion rate, and the disk luminosity increase as the mass input rate increases, $\\propto \\dot{M}_{\\rm input}^{0.7} -\\dot{M}_{\\rm input}^{1.0}$ for $\\alpha=0.5$ and $\\propto \\dot{M}_{\\rm input}^{0.4} -\\dot{M}_{\\rm input}^{0.6}$ for $\\alpha=0.1$. \\subsection{Future Work} As we have already mentioned in \\S 3.1, the sum of the accreting matter and the matter ejected as high-velocity outflows is $20\\%$ of the injected mass, and $80\\%$ of the injected matter is ejected from the computational domain as low-velocity outflows, whose velocities do not exceed the escape velocity at the outer boundary. Since such outflowing matter tends to be accelerated by the radiative force even at the outside of the computational domain, it would be blown away from the system. However, a part of the outflowing matter might return to the vicinity of the black hole through the disk region, since the radiative force does not exceed the gravity near the equatorial plane. Numerical simulations with larger computational domains would make this point clear. Whereas the resulting mass accretion rate onto the black hole is around $10^2 L_{\\rm E}/c^2$ in the present simulations, \\citet{HD07} have indicated that the mass accretion rate can increase up to $10^4 L_{\\rm E}/c^2$. They have reported that the vertical force balance breaks down via a strong radiative force if the mass accretion rate exceeds this limit. However, even in such a case, the matter might accrete onto the black hole, although the strong radiative force would produce powerful outflows. To investigate the maximum value of the accretion rate is an outstanding issue. We should perform numerical simulations with larger computational domains, since the trapping region is expected to expand with the increase of the mass accretion rate. We reveal in the present paper that photons generated deep inside the disk are effectively trapped in the flow, leading to supercritical disk accretion. Although the magnetic fields are not solved in our simulations, magnetic buoyancy might play an important role in the transportation of matter, as well as photons, toward the disk surface (\\citeauthor{Parker75} \\citeyear{Parker75}; \\citeauthor{SR84} \\citeyear{SR84}; \\citeauthor{SC89} \\citeyear{SC89}). Magnetic buoyancy might lead to photon generation near the disk surface if the magnetic fields rise quickly without dissipation deep inside the disk. Thus, magnetic buoyancy would dilute the photon-trapping effect. A photon bubble instability, which is induced in the magnetized, radiation-pressure-dominated region, might also suppress the photon trapping \\citep{Begelman02,Turner05}. In these cases the enhanced radiative force would more effectively accelerate the matter around the disk surface, working to decrease the mass accretion rate. However, the magnetic fields might prevent such acceleration if they strongly tie the matter near the disk surface with the disk matter. In the disk region the matter might easily accrete toward the black hole, since the radiation energy density decreases. Global radiation-magnetohydrodynamic (RMHD) simulations would make these points clear. Local RMHD simulations of accretion flows have been performed by \\citet{Turner03} and \\citet{HKS06}. In addition, it is thought that disk viscosity has magnetic origins (\\citeauthor{HBS01} \\citeyear{HBS01}; \\citeauthor{MMM01} \\citeyear{MMM01}; for a review see \\citeauthor{Balbus03} \\citeyear{Balbus03}). Hence, we should stress again that RMHD simulations are very important to more realistically investigate viscous accretion flows, although an $\\alpha$-viscosity model is employed in the present study." }, "0710/0710.2807_arXiv.txt": { "abstract": "We have identified two moderately hot ($\\sim$18000--22000\\,K) white dwarfs, SDSS\\,J1228+1040 and SDSS\\,J1043+0855, which exhibit double-peaked emission lines in the CaII\\,$\\lambda\\lambda$\\,8600 triplet. These line profiles are unambiguous signatures of gaseous discs with outer radii of $\\sim1R_\\odot$ orbiting the two white dwarfs. Both stars accrete from the circumstellar material, resulting in large photospheric Mg abundances. The absence of hydrogen emission from the discs, and helium absorption in the white dwarf photospheres demonstrates that the circumstellar material is depleted in volatile elements, and the most likely origin of these gaseous rings are tidally disrupted rocky asteroids. The relatively high mass of SDSS\\,J1228+1040 implies that planetary systems can not only form around $4-5\\,M_\\odot$ stars, but may also survive their post main-sequence evolution. ", "introduction": "While more than 250 extra-solar planets orbiting main-sequence stars have been discovered, the destiny of planetary systems in the late stages of the evolution of their host stars is very uncertain, and so far no planet has been found around a white dwarf. Infrared excess detected around a number of white dwarfs has been interpreted as the signature of dust discs \\citep[e.g.][]{zuckerman+becklin87-1, becklinetal05-1, kilicetal06-1, vonhippeletal07-1}. The photospheres of these white dwarfs are rich in metals \\citep{zuckermanetal07-1}, indicating ongoing accretion from the circumstellar material. The likely origin of these debris discs are tidally disrupted asteroids \\citep{jura03-1}, and hence they represent a close association with the planetary systems that the white dwarf progenitor stars may have had. However, while the infrared excess detected around these white dwarfs can be explained in terms of a dusty debris disc, the observations actually do not provide any strong constraint on the geometry of the source of the infrared light \\citep[e.g.][]{reachetal05-1}. We summarise here our recent discovery of two white dwarfs in the SDSS spectroscopic data base which exhibit double-peaked emission lines of Ca\\,II$\\lambda\\lambda$8600, unambiguously confirming a circumstellar disc-like structure \\citep{gaensickeetal06-3, gaensickeetal07-1} ", "conclusions": "It has been suggested that planetary systems may survive the post-main sequence evolution of their host stars \\citep{burleighetal02-1, villaver+livio07-1}, however, no planet has yet been discovered around a white dwarf. The detection of debris discs from rocky asteroids around white dwarfs, such as SDSS\\,J1228+1040 and SDSS\\,J1043+0855 and the cooler white dwarfs with dusty debris discs lends strong support to the survival hypothesis. It appears also entirely possible that these white dwarfs may still have planetesimal objects or planets. SDSS\\,J1228+1040 is particular interesting, as its relatively high mass implies that its progenitor must have had a mass of $\\sim4\\,M_\\odot$ \\citep{dobbieetal06-1}, suggesting that also short-lived massive stars may be host to planetary discs. This is in accordance with the detection of a relatively massive debris disc around the young A2e star MWC\\,480 \\citep{manningsetal97-1}." }, "0710/0710.2813_arXiv.txt": { "abstract": "{ We report on a sensitive search for mid-infrared molecular hydrogen emission from protoplanetary disks. We observed the Herbig Ae/Be stars UX Ori, HD 34282, HD 100453, HD 101412, HD 104237 and HD 142666, and the T Tauri star HD 319139, and searched for H$\\,_2~0-0~S(2)~(J=4-2)$ emission at 12.278 micron and H$\\,_2~0-0~S(1)~(J=3-1)$ emission at 17.035 micron with VISIR, ESO-VLT's high-resolution mid-infrared spectrograph. None of the sources present evidence for molecular hydrogen emission at the wavelengths observed. Stringent 3$\\sigma$ upper limits to the integrated line fluxes and the mass of optically thin warm gas ($T=$ 150, 300 and 1000 K) in the disks are derived. The disks contain less than a few tenths of Jupiter mass of optically thin H$_2$ gas at 150 K, and less than a few Earth masses of optically thin H$_2$ gas at 300 K and higher temperatures. We compare our results to a Chiang and Goldreich (1997, CG97) two-layer disk model of masses 0.02 M$_{\\odot}$ and 0.11 M$_{\\odot}$. The upper limits to the disk's optically thin warm gas mass are smaller than the amount of warm gas in the interior layer of the disk, but they are much larger than the amount of molecular gas expected to be in the surface layer. If the two-layer approximation to the structure of the disk is correct, our non-detections are consistent with the low flux levels expected from the small amount of H$_2$ gas in the surface layer. We present a calculation of the expected thermal H$_2$ emission from optically thick disks, assuming a CG97 disk structure, a gas-to-dust ratio of 100 and T$_{\\rm gas}$ = T$_{\\rm dust}$. We show that the expected H$_2$ thermal emission fluxes from typical disks around Herbig Ae/Be stars are of the order of 10$^{-16}$ to 10$^{-17}$ erg s$^{-1}$ cm$^{-2}$ for a distance of 140 pc. This is much lower than the detection limits of our observations (5 $\\times$ $10^{-15}$ erg s$^{-1}$ cm$^{-2}$). H$_2$ emission levels are very sensitive to departures from the thermal coupling between the molecular gas and dust in the surface layer. Additional sources of heating of gas in the disk's surface layer could have a major impact on the expected H$_2$ disk emission. Our results suggest that in the observed sources the molecular gas and dust in the surface layer have not significantly departed from thermal coupling (T$_{\\rm gas}$/T$_{\\rm dust}<$ 2) and that the gas-to-dust ratio in the surface layer is very likely lower than 1000. ", "introduction": "Circumstellar disks surrounding low- and intermediate-mass stars in their pre-main sequence phase are the locations where planets presumably form. Such protoplanetary disks are composed of gas and dust. Their mass is initially dominated by gas (99\\%), specifically by molecular hydrogen (H$_2$), which is the most abundant gas species. The dust constitutes only a minor fraction of the total disk mass, however, it is the main source of opacity. Consequently, most of what we know observationally about protoplanetary disks has been inferred from studies of dust emission and scattering (for recent reviews see Henning et al. 2006; Natta et al. 2007; Dullemond et al. 2007; Watson et al. 2007). In order to understand the structure and evolution of protoplanetary disks, it is necessary to study their gaseous content independently from the dust. For example, a basic physical quantity such as the disk mass is conventionally deduced from dust continuum emission at millimeter wavelengths assuming an interstellar gas-to-dust ratio of 100 (e.g., Beckwith et al. 1990; Henning et al. 1994). If dust is physically processed in the disk, as should be the case in order to form planets, the gas-to-dust ratio must change with time. The disk dissipation time scale - another fundamental quantity required to disentangle proposed giant planet formation scenarios (Pollack et al. 1996; Boss et al. 1998) - is deduced from observations of thermal infrared excess emission produced by dust grains (Haisch et al. 2001, 2005). Although recent studies (e.g., Sicilia-Aguilar et al. 2006) suggest a parallel evolution of the dusty and gaseous components, it still remains to be demonstrated that the gaseous disks disappear over the same time scale as the infrared excess. A variety of spectral diagnostics of the gas disk have been observed from the UV to the millimeter (see reviews by Najita et al. 2007; Dutrey et al. 2007). However, the only diagnostic that is potentially able to probe the warm gas in the regions where giant planets are thought to form is the mid-infrared (mid-IR) emission lines of H$_2$. UV and near-infrared diagnostics only probe the innermost regions of the disk (R $<$ few AU), and mm and sub-mm diagnostics are limited to probe the cold outermost regions of the disk (R $>$10 AU). \\begin{table*} \\caption{Summary of the observations.} % \\label{table:observations} % \\centering % \\begin{tabular}{@{}l c l c c c c l c c c c c c c @{}} % \\hline\\hline % Star & $\\lambda$ & Date & U.T. & $t_{exp}$ & Airmass $^a$ & V$_{\\oplus\\,{\\rm rad}}\\,^b$ & \\multicolumn{2}{c}{Calibrator $^c$} & $t_{exp}$ & \\multicolumn{2}{c}{Airmass} \\\\ & [$\\mu$m] & & [hh:mm] & [s] & & [km s$^{-1}$] & \\multicolumn{2}{c}{ } & [s] & \\multicolumn{2}{c}{ } \\\\ \\hline % UX Ori & 12.278 & 11 January 2006 & 02:28 & 3600 & 1.0 - 1.2 & 2.06 & HD 36167 & (P)(F) & 1000 & 1.1(P) & 1.2(F) \\\\ & 17.035 & 4 January 2007 & 01:54 & 3600 & 1.0 - 1.0 & 4.50 & HD 25025 & (P) & 600 & 1.0 (P) & ...\\\\ HD 34282 & 12.278 & 10 January 2006 & 04:01 & 3600 & 1.0 - 1.5 & 2.17 & HD 36167 & (P)(F) & 1000 & 1.1(P) & 1.7(F) \\\\ HD 100453 & 12.278 & 22 March 2006 & 07:27 & 3600 & 1.2 - 1.5 & 24.45 & HD 89388 & (P)(F) & 600 & 1.5 (P) & 2.0 (F) \\\\ & 17.035 & 27 March 2006 & 06:26 & 3600 & 1.3 - 1.7 & 22.96 & HD 89388 & (P) & 1000 & 1.5 (P) & ... \\\\ HD 101412 & 12.278 & 30 March 2006 & 04:40 & 3600 & 1.2 - 1.4 & 23.82 & HD 91056 & (P)(F) & 600 & 1.3 (P) & 1.7 (F) \\\\ & 17.035 & 30 March 2006 & 01:27 & 3600 & 1.3 - 1.2 & 24.02 & HD 91056 & (F) & 1000 & ... & 1.3 (F) \\\\ HD 104237 & 12.278 & 10 February 2006 & 05:43 & 3600 & 1.7 - 1.7 & 26.09 & HD 92305 & (P)(F) & 600 & 1.7 (P) & 1.8 (F) \\\\ & 17.035 & 12 February 2006 & 06:36 & 3600 & 1.6 - 1.7 & 26.15 & HD 92305 & (F) & 1000 & ... & 1.8 (F) \\\\ HD 142666 & 17.035 & 28 February 2006 & 06:28 & 3600 & 1.1 - 1.0 & 27.06 & HD 169916 & (P) & 1000 & 1.1 (P) & ...\\\\ HD 319139 & 17.035 & 30 March 2006 & 08:30 & 3600 & 1.1 - 1.0 & 22.47 & HD 169916 & (P) & 1000 & 1.2 (P) & ...\\\\ \\hline \\end{tabular} \\flushleft $^a$ The airmass interval is given from the beginning to the end of the observations. $^b$ V$_{\\oplus\\,{\\rm rad}}$ is the expected velocity shift of the spectra due to the reflex motion of the Earth around the Sun and the radial velocity of the star. $^c$ The standard stars were observed prior (P) and/or immediately following (F) the science observations. \\end{table*} Molecular hydrogen is by far the most abundant molecular species in protoplanetary disks. Unfortunately, H$_2$ is one of the most challenging molecules to detect. Since H$_2$ is a homonuclear molecule, it lacks a permanent dipole moment and its transitions are thus electric quadrupole in nature. The small Einstein coefficients, characteristic of the quadrupole transitions, imply that H$_2$ emission features are very weak. In addition, in the case of protoplanetary disks, the H$_2$ lines are not sensitive to the warm gas in the optically thick regions where the dust and gas are at equal temperature. Practical observational challenges also have to be faced. The mid-IR H$_2$ emission from the disk needs to be detected on the top of a strong mid-IR continuum. From the ground, the mid-infrared windows are strongly affected by sky and instrument background emission, and the H$_2$ transitions at 12 and 17 $\\mu$m lie close to atmospheric absorption lines highly dependent on atmospheric conditions. The advent of high spectral resolution spectrographs mounted on larger aperture telescopes, allows for the first time the study of H$_2$ emission from the ground, but the search is still limited to bright targets. From space, the problems of atmosphere absorption are alleviated and the $J = 2-0$ feature at 28 $\\mu$m is visible. However, the beam sizes are large and the spectral resolution of space mid-infrared spectrographs are usually low when compared to ground-based facilities, therefore, they are not very appropriate for small line-to-continuum ratios. H$_2$ emission from protoplanetary disks in the mid-IR has been reported from ISO observations (Thi et al. 2001). However, subsequent ground-based efforts (Richter et al. 2002; Sheret et al. 2003; Sako et al. 2005) did not confirm the ISO detections. H$_2$ emission in the mid-IR has been searched towards debris disks using Spitzer (Hollenbach et al. 2005, Pascucci et al. 2006, Chen et al. 2006) with no detection reported. Most recently, Bitner et al. (2007) and Martin-Za\\\"idi et al. (2007) reported the detection of mid-IR H$_2$ emission in two Herbig Ae/Be stars (AB Aur and HD 97048) from the ground, and Lahuis et al. (2007) reported the detection of mid-IR H$_2$ emission in 6 T Tauri stars with Spitzer. Here, we report on a sensitive search for molecular hydrogen emission from protoplanetary disks. We observed six southern nearby (d$<$400 pc) Herbig Ae/Be stars and one T Tauri star, employing the Very Large Telescope (VLT) imager and spectrometer for the mid-infrared (VISIR; Lagage et al. 2004)\\footnote{http://www.eso.org/instruments/visir}, ESO's VLT mid-infrared high-resolution spectrograph, and searched for H$_2~0-0~S(1)~(J=3-1)$~emission at 17.035~$\\mu$m and H$_2~0-0~S(2)~(J=4-2)$ emission at 12.278~$\\mu$m. The paper is organized as follows: in Sect. 2 we present the sample studied, and the details of how the observations were performed. In Sect. 3 we discuss the data reduction. In Sect. 4 we deduce upper limits to the H$_2$ fluxes, and using the optically thin approximation, we derive upper limits to the mass of warm (150 - 1000 K) gas in the disks. In Sect. 5 we discuss our results in the context of the Chiang and Goldreich (1997) two-layer disk model. Finally, we summarize our results and conclusions in Sect. 6. \\begin{figure*} \\centering \\includegraphics[angle=0,width=0.95\\textwidth]{7846fig1.eps} \\caption{Spectra obtained for the H$_2$ 0--0 S(2) (J=4--2) line at 12.278 $\\mu$m. The left panel shows a zoom to the -100 to 100 km s$^{-1}$ interval of the atmospheric corrected spectra. A Gaussian of {\\it FWHM} = 15 km s$^{-1}$ and integrated line flux equal to the line-flux upper limits obtained is overplotted at the expected velocity shifted location (vertical dashed line, see Table 2). The central panel shows the full corrected spectra. Dotted lines show spectral regions strongly affected by telluric or standard star absorption features. The right panel shows the continuum normalized spectra of the standard star and the target before telluric correction. The uppermost right panel displays the sky spectrum from a half-chop cycle. The spectra are not corrected for the radial velocity of the targets. } \\end{figure*} \\begin{figure*} \\centering \\includegraphics[angle=0,width=0.95\\textwidth]{7846fig2.eps} \\caption{Spectra obtained for the H$_2$ 0--0 S(1) (J=3--1) line at 17.035 $\\mu$m. The left panel shows a zoom to the -100 to 100 km s$^{-1}$ interval of the atmospheric corrected spectra. A Gaussian of {\\it FWHM} = 15 km s$^{-1}$ and integrated line flux equal to the line-flux upper limits obtained is overplotted at the expected velocity shifted location (vertical dashed line, see Table 2). The central panel shows the full corrected spectra. Dotted lines show spectral regions strongly affected by telluric or standard star absorption features. The right panel shows the continuum normalized spectra of the standard star and the target before telluric correction. The uppermost right panel displays the sky spectrum from a half-chop cycle. The spectra are not corrected for the radial velocity of the targets. } \\end{figure*} ", "conclusions": "We observed a sample of nearby pre-main sequence stars with evidence for cold ($T < 50$ K) gas disk reservoirs and searched for emission of the warm gas with $T > 150$ K, which is expected to be present in the inner region of these disks. None of the targets show any evidence for H$_2$ emission at 17.035 $\\mu$m or at 12.278 $\\mu$m. From the 3$\\sigma$ upper limits of the H$_2$ line fluxes, we found stringent upper limits to the mass of optically thin warm H$_2$. The disks contain less than a few tenths of Jupiter mass of optically thin H$_2$ at 150 K, and less than a few Earth masses of optically thin H$_2$ at 300 K and higher temperatures. Assuming that T$_{\\rm gas}$=T$_{\\rm dust}$ and a gas-to-dust ratio of 100, we compared our results to models of disks employing a Chiang and Goldreich (1997) optically thick two-layer disk model of masses 0.02 M$_{\\odot}$ and 0.11 M$_{\\odot}$. The upper limits to the disk optically thin warm gas mass are smaller than the warm gas mass in the interior layer of the disk, but they are much larger than the amount of molecular gas expected to be in the surface layer. The amount of mass in the surface layer is very small ($<$ 10$^{-2}$ M$_J$) and almost independent of the total disk mass. We calculated the expected H$_2$ S(1) and H$_2$ S(2) line fluxes emitted from a two-layer disk for the low-mass and the high-mass cases assuming a distance of 140 pc, and LTE thermal emission. The predicted line fluxes of the two-layer disk model are of the order of $\\sim$10$^{-16}-10^{-17}$ erg s$^{-1}$ cm$^{-2}$, much smaller than the detection limits of our observations (5 $\\times$ $10^{-15}$ erg s$^{-1}$ cm$^{-2}$). {\\it If the two-layer approximation to the structure of the disk is correct, we are essentially ``blind\" to most of the warm H$_2$ in the disk because it is located in the optically thick interior layer of the disk. Our non-detections are explained because of the small flux levels expected from the little mass of H$_2$ present in the optically thin surface layer.} Naturally, the two-layer disk model is only an approximation of the real structure of a protoplanetary disk. The puffed-up inner rim, which could be an important contributor to the H$_2$ emission, is not included in our models, and in reality there will be a smooth transition zone between the disk hot surface layer and the cool disk interior, which again could contribute significantly to the H$_2$ emission. In addition, the surface layers of the disk are likely to have gas temperatures hotter than the dust temperatures or a gas-to-dust ratio higher than the canonical value of 100. Both effects could potentially increase the H$_2$ emission by significant amounts. We presented additional calculations in which the gas-to-dust ratio and the temperature of the molecular gas in the surface layer of the disk were increased. We showed that detectable S(1) and S(2) H$_2$ line flux levels can be achieved if T$_{\\rm gas\\,surf}$/T$_{\\rm dust\\,surf}>$ 2 and if the gas-to-dust ratio in the surface layer is greater than 1000. H$_2$ emission levels are very sensitive to departures from the thermal coupling between the molecular gas and dust in the surface layer. Our results suggest that in the observed sources the molecular gas and the dust in the surface layer have not significantly departed from thermal coupling and that the gas-to-dust ratio in the surface layer is very likely lower than 1000. A definitive interpretation of our results awaits the development of future, more sophisticated models." }, "0710/0710.2955_arXiv.txt": { "abstract": "We report on the discovery of three new pulsars in the first blind survey of the north Galactic plane (45$^{\\circ}$ < l < 135$^{\\circ}$ ; |b| < 1$^{\\circ}$) with the Giant Meterwave Radio telescope (GMRT) at an intermediate frequency of 610 MHz. The timing parameters, obtained in follow up observations with the Lovell Telescope at Jodrell Bank Observatory and the GMRT, are presented. ", "introduction": "Most pulsar surveys have been carried out with single dish telescopes where there is a trade-off between the collecting area and the beam-width, and consequently the rate of the survey. In a multi-element telescope such as the Giant Meterwave Radio Telescope (GMRT), a large number of smaller antennas can be combined to provide a high sensitivity and yet retain a relatively large beam-width. In this paper, we report on the discovery of three new pulsars in the first blind survey of the north Galactic plane (45$^\\circ$ < l < 135$^\\circ$ ; |b| < 1$^\\circ$) with the GMRT at an intermediate frequency of 610 MHz, which represents the best trade-off between the increased flux density at low frequency for pulsars, interstellar scattering and dispersion and beam-width. The GMRT's multi-element nature was also exploited to determine the positions of the pulsars to an accuracy of 5 arcminutes and this technique is also described. ", "conclusions": "" }, "0710/0710.0999_arXiv.txt": { "abstract": "We investigate the fate of particle production in an expanding universe dominated by a perfect fluid with equation of state $p = \\alpha\\rho$. The rate of particle production, using the Bogolioubov coefficients, are determined exactly for any value of $\\alpha$ in the case of a flat universe. When the strong energy condition is satisfied, the rate of particle production decreases as time goes on, in agreement to the fact that the four-dimensional curvature decreases with the expansion; the opposite occurs when the strong energy condition is violated. In the phantomic case, the rate of particle production diverges in a finite time. This may lead to a backreaction effect, leading to the avoidance of the big rip singularity, specially if $- 1 > \\alpha > - \\frac{5}{3}$. ", "introduction": "The phenomenon of particle production in an expanding universe is due, essentially, to the fact that in curved space-time the vacuum is not unique \\cite{birrel,jacobson}. As the universe evolves, and the curvature changes, the vacuum state also changes: the initial vacuum state, representing the state with no particle, becomes later a multiparticle state. The particle production is directly connected with the curvature of the universe. If the universe is spatially flat, the cosmic evolution leads asymptotically to a Minkowski space-time, where the phenomenon of particle production does not occur anymore. However, this is true only if the strong energy condition is verified: the energy density $\\rho$ and the pressure $p$ must satisfy the relation $\\rho + 3p > 0$. If the strong energy condition is violated the particle production should be zero initially, increasing as the universe evolves, in connection with the increase of the four-dimensional curvature with time when $\\rho + 3p < 0$. \\par In order to compute the particle production in a given cosmological model, it is necessary to fix an initial vacuum state. If the universe is always decelerating, there is no natural and unique choice for this, since all modes are initially outside the Hubble radius feeling strongly the curvature of the space-time. However, a natural choice for the initial vacuum state can be done for the case of an accelerating universe, since all physical modes are initially well inside the Hubble radius, and they behave as in a Minkowski space-time. This is one of the reasons why the primordial inflationary scenario is so successful theoretically: it encodes naturally a mechanism for quantum fluctuations, which will be later the seeds for the large scale structure of the universe \\cite{jerome}. \\par Observations indicate that we live today in an inflationary phase, since the universe is accelerating \\cite{spergel,uzan}. Hence, the mechanism of particle production may play again a very important role today. Suppose we define a quantum state for a given field today. The question we want to address concerns the rate of particle production. In particular, we would like to know if such particle production is so important that it can alters, by back-reaction, the later evolution of the universe. The back-reaction due to quantum effects has already studied in the case of the cosmological constant \\cite{ford2}. The back-reaction may be particularly relevant for the case the universe is dominated by a phantom fluid since, classically, in such a situation the universe will inevitably evolve towards a new singular state, called the big rip. We remember that the observational data favors somehow a phantom scenario today \\cite{caldwell}. A phantom fluid has many special features. Local configurations of a phantom fluid leads to regular black holes \\cite{kirill}. In cosmology, a universe dominated by a phantom fluid implies a future singularity that will occur in a future finite proper time \\cite{frampton}. \\par We will solve the rate of the particle production when the universe is flat and dominated by an equation of state $p = \\alpha\\rho$, for an arbitrary value of $\\alpha$. A simplified scenario, not considering the different phase transitions that occur in the real universe, will be adopted. The calculations will be performed for a massless scalar field, for which the corresponding Klein-Gordon equation is solved. Through the quantization of the field, the rate of particle may be determined using the technique of Bogolioubov's coefficients. The general solution for this problem is expressed in terms of Bessel functions. The rate of particle production is determined exactly for any value of $\\alpha$. The connection of this rate with the strong energy condition is established. It is shown that for a phantom fluid the rate of particle production diverges as the big rip is approached. This may lead to a back-reaction effect changing the effective equation of state of the universe, implying perhaps the avoidance of the big rip itself: the particle produced does not necessarily obeys the phantom equation of state and may become the dominant component of the universe if $ - 1 > \\alpha > - \\frac{5}{3}$. Hence, the back-reaction of the particle production may be an ineluctable mechanism to the avoidance of the big rip if the pressure is not excessively negative. The limiting case, $\\alpha = - \\frac{5}{3}$, has already been found before in a complete different context, that is, in the evolution of classical scalar perturbations in phantom cosmological models \\cite{finelli,fabris}. \\par This paper is organized as follows. In the next section we review the formalism of particle creation in a curved space-time. In section $3$ we apply the formalism to a perfect fluid cosmological model, and we determine the rate of particle creation for any value of $\\alpha$. In section $4$ we present our conclusions. Even if many steps of the formalism used and of the calculation performed are quite well known, we will present them with some details in order to trace the main features of the final results. ", "conclusions": "We have evaluate the rate of creation of massless scalar particle in a universe dominated by a perfect fluid whose equation of state is given by $p = \\alpha\\rho$. An analytic expression has been found in terms of Hankel's functions. Since we have not considered a complete cosmological scenario, with a sequence of different phases, with an initial inflationary phase (as in standard cosmological model), the calculation performed here makes sense, strictly speaking, only when the strong energy condition is violated, that is, $\\alpha < - 1/3$. For such a case, there is a natural initial vacuum state, from which the particle occupation number can be determined. However, we formally extended the calculation for any value of $\\alpha$. It has been found that for those \"dark energy\" scenarios the rate of particle creation diverges as $t \\rightarrow \\infty$, that is, in the future infinity. \\par We have payed special attention to the phantom scenario. The reason is that a universe dominated by a phantom fluid may develop a singularity in a finite future time, the so-called big rip. In this case, we have found that it is possible that the energy density associated to the particles created (which obey in present case an equation of state of the type $p_s = \\rho_s$) becomes dominant over the phantom fluid if $- 1 > \\alpha > - 5/3$. Hence, the big rip can be avoided if the pressure is not deeply negative. \\par It is interesting to remark that a similar critical point, $\\alpha = - 5/3$, has been found in the case of classical scalar perturbation \\cite{fabris}. In that case, however, the evolution of scalar perturbation may destroy the conditions of homogeneity (necessary for the big rip) if $\\alpha < - 5/3$, that is, if the pressure is negative enough. The result found here is exactly the opposite: quantum effects can be operative in the sense of destroying the conditions for the big rip if the pressure is not negative enough. It must be stressed, however, that the evaluation made in the present work must be complemented by a study of more general quantum fields and by a deeper thermodynamical analysis of the energy balance between the phantom fluid and the created particles as the big rip is approached. \\newline \\vspace{0.5cm} \\newline {\\bf Acknowledgements}:\\\\ We thank CNPq (Brazil) and the french/brazilian scientific cooperation CAPES/COFECUB (project number 506/05) for partial financial support. We thank specially J\\'er\\^ome Martin for his criticisms on the text." }, "0710/0710.1167_arXiv.txt": { "abstract": "A symplectic integrator algorithm suitable for hierarchical triple systems is formulated and tested. The positions of the stars are followed in hierarchical Jacobi coordinates, whilst the planets are referenced purely to their primary. The algorithm is fast, accurate and easily generalised to incorporate collisions. There are five distinct cases -- circumtriple orbits, circumbinary orbits and circumstellar orbits around each of the stars in the hierarchical triple -- which require a different formulation of the symplectic integration algorithm. As an application, a survey of the stability zones for planets in hierarchical triples is presented, with the case of a single planet orbiting the inner binary considered in detail. Fits to the inner and outer edges of the stability zone are computed. Considering the hierarchical triple as two decoupled binary systems, the earlier work of Holman \\& Wiegert on binaries is shown to be applicable to triples, except in the cases of high eccentricities and close or massive stars. Application to triple stars with good data in the multiple star catalogue suggests that more than 50 \\% are unable to support circumbinary planets, as the stable zone is non-existent or very narrow. ", "introduction": "\\label{sec:intro} It is well known that many of the stars in the solar neighbourhood exist in multiple systems. As the number of planetary surveys increases, planets are regularly being found not only in single star systems, but binaries and triples as well. For example, recently a hot Jupiter has been claimed in the triple system HD 188753 (Konacki 2005; but see also Eggenberger et al. 2007), and \\citet{DB07} lists 33 binary systems and 2 other hierarchical triples known to harbour exoplanets. As the majority of work on planetary dynamics has been for single star systems, the dynamics of bodies in these multiple stellar systems is of great interest. At first sight, it might not seem likely to expect long term stable planetary systems to exist in binary star systems, let alone triples, but numerical and observational work is starting to show otherwise. In recent years, much study has been devoted to the stability of planets and planetesimal discs in binary star systems. There are several investigations of individual systems (e.g. work on $\\gamma$ Cephei by \\citealt{Dv03}, \\citealt{Ha06} and \\citealt{VE06}) as well as substantial amounts of work on more general limits for stability of smaller bodies in these systems. Notably, \\citet{HW99} approach this problem by using numerical simulation data to empirical fit critical semimajor axes for test particle stability as functions of binary mass ratio and eccentricity. These general studies are of great use in the investigation of observed systems and their stability properties, giving an effective and fast method of placing limits on stability in the large parameter space created by observational uncertainties. The aim of this work is to investigate test particle stability in hierarchical triple star systems, and to see if any similar boundaries can be defined for them. To do this, a new method for numerically integrating planets in such systems is presented, following the ideas for binary systems presented by \\citet{CQDL02}. Although there have been empirical studies of the stability of hierarchical star systems themselves, there do not appear to have been studies of small bodies in such systems. There is a great deal of literature concerning periodic orbits in the general three and four body problems (see e.g. \\citealt{Sz67}), but these are almost entirely devoted to considerations of planetary satellites in single star systems, for example satellite transfer in the Sun-Earth-Moon systems. Also, while periodic orbits prove that stable solutions exist in these problems, they are of little practical use in determining general stability limits. The problem of planetary orbits in triple systems is more complex than for those in binary systems, as there are many different orbital configurations possible relative to the three stars. These are considered in Section~\\ref{sec:orbits}, and a method for classifying them is described in order to simplify the discussion in this paper. The derivation of a method to numerically integrate such planets is then given in Section~\\ref{sec:maths}. In Section~\\ref{sec:stats} is a brief overview of the statistical properties of known triple systems, as a basis for deciding the parameters of the systems used in the numerical simulations. The results of numerical investigations into stability properties are then presented in Section~\\ref{sec:sp} for one of the possible orbital types. ", "conclusions": "\\label{sec:conclusion} The main achievement of this paper is the formulation of a symplectic integrator algorithm suitable for hierarchical triple systems. This extends the algorithm for binary systems presented by \\citet{CQDL02}. The positions of the stars are followed in hierarchical Jacobi coordinates, whilst the planets are referenced purely to their primary. Each of the five distinct cases, namely circumtriple orbits, circumbinary orbits and circumstellar orbits around each of the stars in the hierarchical triple, requires a different splitting of the Hamiltonian and hence a different formulation of the symplectic integration algorithm. Here, we have given the mathematical details for each of the five cases, and presented a working code that implements the algorithm. As an application, a survey of the stability zones for circumbinary planets in hierarchical triples is presented. Here, the planet orbits an inner binary, with a more distant companion star completing the stellar triple. Using a set of numerical simulations, we found the extent of the stable zone which can support long-lived planetary orbits and provided fits to the inner and outer edges. The effect of low inclination on this boundary is minimal. A reasonable first approximation to a behaviour of a hierarchical triple is to regard it as a superposition of the dynamics of the inner binary and a pseudo-binary consisting of the outer star and a point mass approximation to the inner binary. If it is considered as two decoupled binary systems, then the earlier work of Holman \\& Wiegert (1999) on binaries is applicable to triples, except in the cases of high eccentricities and close or massive stars. The implication here is that the addition of a stable third star does not distort the original binary stability boundaries. As mentioned, \\citet{MW06} have shown that overlapping sub-resonances are the cause of the boundary in the binary case. It is reasonable to expect that in triples the same process is responsible, and the similarities between the binary and triple results support this theory. It is also expected that there is a regime in which the resonances from each sub-binary start to overlap as well, further destabilising the test particles. Evidence of this is the deviation from the binary results when the stars are close, massive and very eccentric, when resonances would be both stronger and wider. Since the parameter space investigated was chosen to reflect the observed systems it would seem to be a reflection on the characteristics of known triple stars that most lie in the decoupled regime. The relatively constant nature of the stellar orbits in the simulations is however a consequence of the test particle orbits being destabilised long before the stars are close enough to interact. By extension of all these arguments, it is expected that the binary criteria can be used to fairly accurately predict the stability zones in any hierarchical stellar system, no matter the number of stars. The results presented here can be used to estimate the number of known hierarchical triple systems that could harbour S(AB)-P planets. \\citet{To97} lists 54 systems with semimajor axis, eccentricity and masses for both the inner and outer components. The mutual inclinations of most are not well known, but there are nine systems listed for which this angle can be determined. For five of these it is less than $15^\\circ$, two are around $40^\\circ$ and two are retrograde. Although a small sample it suggests that there are systems that fall within the low inclination regime investigated here. Using the criteria of \\cite{HW99} and those found here for the position of the inner and outer critical semimajor axis the size of the coplanar stable region for each of these triples can be calculated. This can be considered an upper limit, since it is likely that very non-coplanar systems and those with significant eccentricities for both binary components will further destabilise planets. Of the 54 systems 13 are completely unstable to circumbinary planets according to the four parameter fits (compared to 11 using \\citet{HW99}'s criteria). Figure \\ref{fig:zone} shows a plot of the width of the stable region for the remaining systems. Interestingly, the majority seem to have very small stable zones, with 16 smaller than an au. Whether this is a feature of triple systems, or an observation bias is not apparent. It does indicate though that circumbinary planets are unlikely to exist in at least 50 \\% of observable systems. \\begin{figure} \\centering \\includegraphics[width=\\textwidth]{fig14.eps} \\caption{ Widths of the circumbinary stability zones for known triple systems in au and as a percentage of the area between the two sub-binaries, calculated using the four parameter fits derived in Section~\\ref{sec:sp}. Note that using the criteria of \\citet{HW99} gives almost identical results.} \\label{fig:zone} \\end{figure}" }, "0710/0710.1021_arXiv.txt": { "abstract": "\\noindent We test statistically the hypothesis that radio pulsar glitches result from an avalanche process, in which angular momentum is transferred erratically from the flywheel-like superfluid in the star to the slowly decelerating, solid crust via spatially connected chains of local, impulsive, threshold-activated events, so that the system fluctuates around a self-organised critical state. Analysis of the glitch population (currently 285 events from 101 pulsars) demonstrates that the size distribution in individual pulsars is consistent with being scale invariant, as expected for an avalanche process. The measured power-law exponents fall in the range $-0.13\\leq a \\leq 2.4$, with $a\\approx 1.2$ for the youngest pulsars. The waiting-time distribution is consistent with being exponential in seven out of nine pulsars where it can be measured reliably, after adjusting for observational limits on the minimum waiting time, as for a constant-rate Poisson process. PSR J0537$-$6910 and PSR J0835$-$4510 are the exceptions; their waiting-time distributions show evidence of quasiperiodicity. In each object, stationarity requires that the rate $\\lambda$ equals $- \\epsilon \\dot{\\nu} / \\langle\\Delta\\nu\\rangle$, where $\\dot{\\nu}$ is the angular acceleration of the crust, $\\langle\\Delta\\nu\\rangle$ is the mean glitch size, and $\\epsilon\\dot{\\nu}$ is the relative angular acceleration of the crust and superfluid. Measurements yield $\\epsilon \\leq 7 \\times 10^{-5}$ for PSR J0358$+$5413 and $\\epsilon \\leq 1$ (trivially) for the other eight objects, which have $a < 2$. There is no evidence that $\\lambda$ changes monotonically with spin-down age. The rate distribution itself is fitted reasonably well by an exponential for $\\lambda \\geq 0.25\\,{\\rm yr^{-1}}$, with $\\langle \\lambda \\rangle = 1.3^{+0.7}_{-0.6}\\,{\\rm yr^{-1}}$. For $\\lambda < 0.25\\,{\\rm yr^{-1}}$, its exact form is unknown; the exponential overestimates the number of glitching pulsars observed at low $\\lambda$, where the limited total observation time exercises a selection bias. In order to reproduce the aggregate waiting-time distribution of the glitch population as a whole, the fraction of pulsars with $\\lambda > 0.25\\,{\\rm yr^{-1}}$ must exceed $\\sim 70$ per cent. ", "introduction": "} Glitches are tiny, impulsive, randomly timed increases in the spin frequency $\\nu$ of a rotation-powered pulsar, sometimes accompanied by an impulsive change in the frequency derivative $\\dot{\\nu}$. They are to be distinguished from timing noise, a type of rotational irregularity where pulse arrival times wander continuously, although there is evidence that timing noise is the cumulative result of frequent microglitches in certain pulsars \\citep{cor85,dal95}. At the time of writing, 285 glitches in total have been detected in 101 objects ($\\sim 6\\%$ of the known radio pulsar population), the majority in the last four years, facilitated by the Parkes Multibeam Survey, refined multifrequency ephemerides, and better interference rejection algorithms \\citep{hob02,kra03,kra05,lew05,jan06}. Efforts to analyse the data statistically have focused on the correlation of glitch activity with age \\citep{mck90,she96,ura99,lyn00,wan00} and Reynolds number \\citep{peralta06,mel07}, the post-glitch relaxation time-scale \\citep{wan00,won01}, the size distribution \\citep{mor93a,mor93b,peralta06}, and the correlation between glitch sizes and waiting times \\citep{wan00,won01,mid06,peralta06}. \\citet{hob02} reviewed the role of observational selection effects. Most glitching pulsars ($65\\%$) have been seen to glitch once, but a minority glitch repeatedly; the current record holder is PSR J1740$-$3015, with 33 glitches. Of those objects which glitch repeatedly, most do so at unpredictable intervals, but two (PSR J0537$-$6910 and Vela) are quasiperiodic; Vela, in particular, has been likened to a relaxation oscillator \\citep{lyn96}. The fractional increase in $\\nu$ spans seven decades ($3\\times 10^{-11} \\leq \\Delta\\nu \\leq 2\\times 10^{-4}$) across the glitch population and as many as four decades in a single object (e.g.\\ $7\\times 10^{-10} \\leq \\Delta\\nu \\leq 2\\times 10^{-6}$ in PSR J1740$-$3015). The spin-down age $\\tau_{\\rm c}= -\\nu/(2\\dot{\\nu})$ of glitching pulsars spans four decades, from $1\\times 10^3\\,{\\rm yr}$ to $3\\times 10^7\\,{\\rm yr}$. In many respects, therefore, the glitch phenomenon is {\\em scale invariant}. This striking property invites physical interpretation. Theories of pulsar glitches have focused mainly on the local microphysics of the superfluid in the stellar interior and its coupling to the solid crust, for example the strength of vortex pinning \\citep{and75,jon98}, the rate of vortex creep \\citep{lin96}, or the conditions for exciting superfluid turbulence \\citep{per05,per06,mel07,and07}. Ultimately, however, the local microphysics must be synthesized with the global, {\\em collective} dynamics in order to make full contact with observational data. (Likewise, a practical model of earthquakes must synthesize the microphysics of rock fracture with the macrodynamics of interacting tectonic plates.) For example, if approximately $10^{16} (\\Delta\\nu/1\\,{\\rm Hz})$ vortices unpin from crustal lattice sites in sympathy during a glitch, they must communicate rapidly across distances much greater than their separation. How? And why does the number that unpin fluctuate so dramatically (by up to four orders of magnitude) from glitch to glitch in a single pulsar, while always amounting to a small fraction ($\\Delta\\nu/\\nu$) of the total? Such collective, scale invariant behavior is a generic feature of a class of natural and synthetic far-from-equilibrium systems, called self-organized critical systems, that are discrete, interaction dominated, and slowly driven, and that adjust internally via erratic, spatially connected {\\em avalanches} of local, impulsive, threshold-activated, relaxation events \\citep{jen98}. Such systems fluctuate around a stationary state towards which they evolve spontaneously, in which global driving balances local relaxation on average over the long term. The archetype of a self-organized critical system is the sandpile \\citep{bak87}. In this paper, we study pulsar glitches as an avalanche process, as first proposed by \\citet{mor93a}. After reviewing self-organised criticality in \\S\\ref{sec:gli2}, we define the statistical sample on which our study is based (\\S\\ref{sec:gli3}) and analyze the observed distribution of glitch sizes (\\S\\ref{sec:gli4}) and waiting times (\\S\\ref{sec:gli5}). Some implications for glitch physics are explored in \\S\\ref{sec:gli6}. We only include radio pulsars in the sample, to preserve its homogeneity, even though glitches have now been observed in anomalous X-ray pulsars (magnetars) as well \\citep{dal03,kas03}. ", "conclusions": "} In this paper, we analyze the size and waiting-time distributions of pulsar glitches, taking advantage of the enlarged data set produced by the Parkes Multibeam Survey. We conclude that the data are consistent with the hypothesis that pulsar glitches arise from an avalanche process. In each of seven pulsars with $N_{\\rm g} > 5$, the size distribution is consistent with being scale invariant across the observed range of $\\Delta\\nu$ (up to four decades), and the waiting-time distribution is consistent with being Poissonian. These features are natural if the system is driven globally at a constant rate (as the pulsar spins down), and each glitch corresponds to a locally collective, threshold activated relaxation of one of the many spatially independent, metastable stress reservoirs in the system (e.g. via a vortex unpinning or crust cracking avalanche). In two pulsars, PSR J0537$-$6910 and PSR J0835$-$4510, the dynamics may include a second, quasiperiodic component, comprising $\\sim 20\\%$ of the events. The size and waiting-time distributions of the quasiperiodic component are narrowly peaked, as expected for rare, system-spanning avalanches, which relax a large fraction of the total stress accumulated in the system. This two-component behavior is observed widely in self-organised critical systems, including experiments on magnetic flux vortices in type II superconductors, which are closely analogous to neutron star superfluids \\citep{fie95}. The power-law exponent of the size probability density function differs from pulsar to pulsar, spanning the range $-0.13 \\leq a \\leq 2.4$. Calculating $a$ theoretically from first principles is a deep problem which has not yet been solved for other self-organised critical systems, let alone glitching pulsars, although some progress has been made on two-dimensional sandpiles using renormalization group techniques \\citep{pie94,jen98}. In the mean-field approximation, which is exact in four or more dimensions, theoretical calculations on sandpiles (and other systems in their universality class) yield $a = 1.5$, whereas three-dimensional cellular automata output $a=1.3$ \\citep{jen98}. The size distribution transmits two important lessons concerning the microphysics of glitches. First, the fact that $a$ differs from pulsar to pulsar implies that the strength and level of conservation of the local (e.g.\\ pinning and intervortex) forces also differs \\citep{ola92,fie95}. By contrast, in equilibrium critical systems like ferromagnets, $a$ depends only on the dimensionality of the system and its order parameter and is therefore universal. Second, except for the two pulsars which show quasiperiodicity, $a$ appears to vary smoothly with spin-down age, with $a\\approx 1.2$ for the youngest pulsars (e.g.\\ the Crab). Figure \\ref{fig:gli9} depicts the trend between $a$ and $\\tau_{\\rm c}$. It is suggestive; after all, local pinning forces do depend on temperature and hence $\\tau_{\\rm c}$. Interestingly, however, there is no clear trend between $a$ and $\\nu$, even though the mean vortex spacing (and hence intervortex force) is proportional to $\\nu^{1/2}$. It will pay to study these trends more thoroughly as more glitch data is collected. An avalanche process predicts a specific relation between the distributions of glitch sizes $\\Delta\\nu$ and lifetimes $T$ (as opposed to waiting times $\\Delta t$). Specifically, in a self-organized critical state, the lifetime probability density function is also a power law, $p(T) \\propto T^{-b}$, with \\begin{equation} b = 1 + (a-1)\\gamma_2 / \\gamma_3~. \\label{eq:gli16} \\end{equation} The constants $\\gamma_2$ and $\\gamma_3$ are defined such that the cardinality of an avalanche scales with its linear extent ($L$) as $L^{\\gamma_2}$ and its lifetime (i.e.\\ duration) scales as $L^{\\gamma_3}$ \\citep{jen98}. Both $\\gamma_2$ and $\\gamma_3$ depend on the effective dimensionality of the local forces and can be calculated numerically using a cellular automaton. In two dimensions, avalanches are compact, not fractal, and one has $\\gamma_2=2$; in three dimensions, one has $2 < \\gamma_2 < 3$. At present, radio timing experiments cannot measure $T$; most glitches are detected as unresolved, discontinuous, spin-up events with $T < 120\\,{\\rm s}$ \\citep{mcc90}. \\footnote{In the Crab, some spin-up events seem to be resolved, e.g.\\ at epochs MJD 50260 ($T \\approx 0.5\\,{\\rm d}$) and MJD 50489 (secondary spin up, $T \\approx 2\\,{\\rm d}$) \\citep{won01}. If these are rare but otherwise standard glitches originating from the long-$T$ tail of the lifetime distribution, it is puzzling that other, shorter, but still resolved (and presumably more common) spin-up events are not observed, e.g.\\ with $T\\sim 0.1\\,{\\rm d}$ or $0.01\\,{\\rm d}$. Alternatively, the events at MJD 50620 and MJD 50489 may have been triggered by a different physical mechanism.} In the future, however, single- and/or giant-pulse timing experiments with more sensitive instruments (e.g.\\ the Square Kilometer Array) will test this prediction. If confirmed, it will independently corroborate the avalanche hypothesis. The mean glitching rates of the nine pulsars studied here are fairly narrowly distributed, spanning the range $0.35\\,{\\rm yr^{-1}} \\leq \\lambda \\leq 2.6\\,{\\rm yr^{-1}}$. The probability density function for $\\lambda$ is adequately fitted by an exponential, as for solar flare avalanches \\citep{whe00}, with $\\langle\\lambda\\rangle = 1.3_{-0.6}^{+0.7}$ {\\rm yr}$^{-1}$, or by an exponential with a lower cutoff, at $\\lambda_{\\rm min} \\approx 0.25$ {\\rm yr}$^{-1}$. A theoretical derivation of $\\langle\\lambda\\rangle$ from first principles is currently lacking, although estimates of how long it takes to crack the crust locally predict reasonable rates, if the critical strain angle approaches that of imperfect terrestrial metals \\citep{alp96,mid06}. Figure \\ref{fig:gli10} plots $\\lambda$ versus $\\tau_c$ for the nine pulsars examined individually in this paper. There is no significant trend. The data are consistent with the notion that old pulsars glitch less frequently than young pulsars \\citep{she96}, but they are equally consistent with the notion that the glitching rate is independent of age. Many authors have searched for a correlation between waiting time and the size of the next glitch. Such a correlation appears to be absent from the data, e.g.\\ Figure 17 in \\citet{wan00} and Figure 10 in \\citet{won01}. At first blush, this is surprising: the vortex unpinning and crust fracture paradigms, which are driven by the accumulation of differential rotation and mechanical stress respectively, seem to be natural candidates for a `reservoir effect'. Avalanche dynamics resolves this apparent paradox. The reservoir effect does operate locally, but the star contains many reservoirs, insulated from each other by relaxed zones, whose storage capacities evolve stochastically in response to the slow driver and avalanche history. During a glitch, a single reservoir (often small but sometimes large) relaxes at random via an avalanche, releasing its stored $\\Delta\\nu$ (and destabilizing neighboring reservoirs in preparation for the next glitch). Some of the $\\Delta\\nu$ is accumulated since the previous glitch, but the remainder is `borrowed' from earlier epochs, when some other reservoir relaxed instead. All self-organized critical systems share these dynamics; the waiting time is uncorrelated with the size of the next avalanche \\citep{jen98}. The only exceptions are large, system-spanning avalanches, which always have roughly the same sizes and waiting times, and which account for $\\sim 20\\%$ of the glitches in PSR J0537$-$6910 and PSR J0835$-$4510. A corollary of the previous paragraph is that the total $\\Delta\\nu$ released in glitches up to some epoch is less than the total crust-superfluid differential rotation accumulated since that epoch, viz. \\begin{equation} \\sum_{i=1}^{N_{\\rm g}} \\Delta\\nu_i \\leq \\epsilon |\\dot{\\nu}| \\sum_{i=1}^{N_{\\rm g}} \\Delta t_{i}, \\label{eq:gli17} \\end{equation} where $\\epsilon\\dot{\\nu}$ is the relative angular acceleration of the crust and superfluid due to electromagnetic spin down. The `staircase' described by (\\ref{eq:gli17}) has been noted previously \\citep{she96,lyn00}, both in quasiperiodic glitchers like PSR J0537$-$6910 [e.g.\\ Figure 8 in \\citet{mid06}], where the reservoir effect is obvious, and in Poisson glitchers like PSR J0534$+$2200, [e.g.\\ Figure 12 in \\citet{won01}], where the trend is more subtle because it reverts to the mean over long times, not after every glitch. Upon dividing (\\ref{eq:gli17}) by $N_{\\rm g}$, and averaging over long times, the inequality becomes an equality (provided there is no secular accumulation of differential rotation in the system) and we recover (\\ref{eq:gli5}). It is fundamentally impossible to measure $\\epsilon$ for individual pulsars with current data, because $\\langle \\Delta \\nu \\rangle$ is dominated by large (and therefore rare) glitches for $a < 2$. It is therefore wrong to assume stationarity over a typical, $40$-yr observation interval. Consequently, we are prompted to reassess the familiar correlation between activity and spin-down age \\citep{she96}. Our definition of $\\epsilon\\dot{\\nu}$ is identical to $\\dot{\\nu}_{\\rm glitch}$ in \\citet{lyn00} (but for individual objects, not in aggregate) and $A_{\\rm g}$ in \\citet{won01}. It is closely related to the original activity parameter defined by \\citet{mck90}, which equals $N_{\\rm g} \\nu^{-1} \\epsilon |\\dot{\\nu}|$. For PSR J0358$+$5413, we measure $\\epsilon\\leq 7 \\times 10^{-5}$, lower than the {\\it aggregate} value $0.017\\pm0.002$ measured by \\citet{lyn00} for objects with $\\tau_{\\rm c} > 10\\,{\\rm kyr}$ (binned by semi-decades in $\\dot{\\nu}$). \\footnote{The aggregate value $\\dot{\\nu}_{\\rm glitch}$ \\citep{lyn00}, binned over semi-decades in $\\dot{\\nu}$, effectively averages together different pulsars. While this approach reduces the formal error bar on $\\dot{\\nu}_{\\rm glitch}$, its physical interpretation is less straightforward, given the likelihood that $\\epsilon$ is different in different pulsars.} Interpreted in terms of the vortex unpinning model, this result suggests that $0.007$--$2$ \\% of the angular momentum outflow during spin down may be stored in metastable reservoirs on average over time. On the other hand, five other objects have $0.04 \\leq \\epsilon_{\\rm max} \\leq 0.8$, under the questionable assumption that the maximum physical size is $\\Delta \\nu_{\\rm upper} = 2 \\times 10^{-4}$ in all pulsars. Our data are therefore inadequate to update usefully the value $A_{\\rm g}/|\\dot{\\nu}| = 1\\times 10^{-5}$ measured by \\citet{won01} for PSR J0534$+$2200. In the context of vortex unpinning, it has been argued that the aggregated $\\epsilon$ measured by \\citet{lyn00} partly corroborates the hypothesis that younger pulsars are still in the process of forming their capacitive elements, e.g.\\ by creating pinning centers through crust fracture, while older pulsars have mostly completed the task \\citep{alp96,won01}. However, the full picture is more complicated. Vela's quasiperiodic avalanches point to a richly connected network of reservoirs \\citep{alp96}, yet its aggregated value $\\dot{\\nu}_{\\rm glitch}$ is relatively low. On the other hand, the other quasiperiodic glitcher, PSR J0537$-$6910, is relatively young ($\\tau_{\\rm c} = 4.9\\,{\\rm kyr}$); how did it form a richly connected reservoir network so quickly? And, if its network is so richly connected, why is its aggregated $\\dot{\\nu}_{\\rm glitch}$ value so low? Likewise, PSR J0358$+$5413 is the oldest object in the sample ($\\tau_{\\rm c} = 560 \\,{\\rm kyr}$), yet its $\\epsilon$ value arguably points to a dearth of capacitive elements, characteristic of a young object. There are no obvious grounds (e.g.\\ quasiperiodicity) on which to treat PSR J0358$+$5413 as exceptional. Do all pulsars glitch eventually? It has been speculated in the past that there is something special physically about the minority of pulsars that do glitch. While it is impossible to reject this hypothesis unequivocally with the data at hand, the results presented here suggest that all pulsars are capable of glitching. However, most do so infrequently (low $\\lambda$) and hence have not been detected during the last four decades of timing experiments. We find that up to $\\sim 30 \\%$ of the pulsar population can glitch at rates lower than $\\lambda_{\\rm min}= 0.25$ {\\rm yr}$^{-1}$ and still conform with the measured aggregate waiting-time distribution. Once verified, the claimed Poissonian nature of the glitch mechanism can be invoked to exclude broad classes of glitch theories, e.g.\\ those that rely on `defects' or `turbulence' at special locations (like the pole), or that involve a pair of dependent events (A.\\ Martin, private communication). It is important to interpret aftershocks carefully in this light \\citep{won01}. In self-organized critical systems, the excess number of avalanches following a large avalanche (over and above the Poissonian baseline following a small avalanche) scales inversely with the time elapsed, a property known as Omori's law for earthquakes \\citep{jen98}. In this paper, we do not analyze post-glitch relaxation times and glitch-activated changes in $\\dot{\\nu}$ in the context of avalanche processes, e.g.\\ the correlation between $\\Delta\\dot{\\nu}$ and the transient component of $\\Delta\\nu$ \\citep{won01}. We also assume implicitly that the quantized superfluid vortices in the vortex unpinning model are organized in a rectilinear array, even though recent work suggests that meridional circulation destabilizes the array and converts it into a turbulent tangle \\citep{per05,per06}. Further study of these matters is deferred to future work. We acknowledge the computer time and system support supplied by the Australian Partnership for Advanced Computation (APAC) and the Victorian Partnership for Advanced Computation (VPAC). We thank Andre Trosky for illuminating conversations on self-organized criticality and cellular automata. This research was supported by a postgraduate scholarship from the University of Melbourne. It makes use of the Australia Telescope National Facility Pulsar Catalogue \\citep{man05}, which can be accessed on-line at {\\tt http://www.atnf.csiro.au/research/pulsar/psrcat}." }, "0710/0710.3815_arXiv.txt": { "abstract": "The Newtonian solid-mechanical theory of non-compressional spheroidal and torsional nodeless elastic vibrations in the homogenous crust model of a quaking neutron star is developed and applied to the modal classification of the quasi-periodic oscillations (QPOs) of X-ray luminosity in the aftermath of giant flares in SGR 1806-20 and SGR 1900+14. A brief outline is given of Rayleigh's energy method which is particular efficient when computing the frequency of nodeless elastic spheroidal and torsional shear modes as a function of multipole degree of nodeless vibrations and two input parameters -- the natural frequency unit of shear vibrations carrying information about equation of state and fractional depth of peripheral seismogenic layer. In so doing we discover that the dipole overtones of both spheroidal and torsional nodeless vibrations possess the properties of Goldstone soft modes. It is shown that obtained spectral formulae reproduce the early suggested identification of the low-frequency QPOs from the range $30\\leq \\nu \\leq 200$ Hz as torsional nodeless vibration modes $\\nu(_0t_\\ell)$ of multipole degree $\\ell$ in the interval $2\\leq \\ell \\leq 12$. Based on this identification, which is used to fix the above mentioned input parameters of derived spectral formulae, we compute the frequency spectrum of nodeless spheroidal elastic vibrations $\\nu(_0s_\\ell)$. Particular attention is given to the low-frequency QPOs in the data for SGR 1806-20 whose physical origin has been called into question. Our calculations suggest that these unspecified QPOs are due to nodeless dipole torsional and dipole spheroidal elastic shear vibrations: $\\nu(_0t_1)=18$ Hz and $\\nu(_0s_1)=26$ Hz. ", "introduction": "The discovery of QPOs of X-ray luminosity in the aftermath of giant flares SGR 1806-20 and SGR 1900+14 (Israel et al 2005; Strohmayer \\& Watts 2006; Israel 2007; Watts \\& Strohmayer 2007) with concomitant interpretation of QPOs as caused by quake-induced differentially rotational seismic vibrations has stimulated remarkable developments in the magnetar asteroseismology (e.g. Piro 2005, Samuelsson \\& Andersson 2007a, 2007b; Lee 2007; Levin 2007; Watts \\& Reddy 2007; Sotani et al 2007; Bastrukov et al 2007a and references therein). Following the above interpretation and presuming the dominant role of the elastic restoring force, the focus of most theoretical works is on computing the frequency spectra of odd-parity torsional mode of shear vibrations and less attention is paid to the even parity spheroidal elastic mode. However, from the viewpoint of modern global seismology (Lay \\& Wallace 1995; Aki \\& Richards 2003), the spheroidal vibrational mode in a solid star and planet has the same physical significance as the toroidal one in the sense that these two fundamental modes owe their existence to one and the same restoring force (e.g. McDermott, Van Horn, Hansen 1988; Bastrukov, Weber, Podgainy 1999, Bastrukov et al 2007b). In this light there is a possibility that, by not considering both these modes on an equal footing, we may miss discovering certain essential novelties which are consequences of solid mechanical laws governing seismic vibrations of superdense matter of neutron stars. Adhering to this attitude and continuing the investigations recently reported by Bastrukov et al (2007a), we derive here spectral equations for the frequency of both spheroidal and torsional elastic nodeless vibrations in the solid crust of quaking neutron star and examine what conclusions can be drawn regarding low-frequency QPOs whose physical nature still remain unclear. In Sec.2 by use of the energy variational method we derive spectral formulae for the frequencies of the nodeless spheroidal and torsional elastic vibrations locked in the finite-depth seismogenic layer. Particular attention is given to the dipole spheroidal and torsional vibrations possessing properties of Goldstone's soft modes. In sec.3, the obtained spectral formulae are applied to a modal analysis of available data on the above mentioned QPOs. The obtained results are highlighted in Sec.4. ", "conclusions": "The exact spectral formulae, which has been obtained here within the framework of Newtonian, non-relativistic, solid-mechanical theory of seismic vibration for the first time, are interesting in its own right from the viewpoint of general theoretical seismology (e.g. Lay, Wallace 1995) because they can be utilized in the study of seismic vibrations of more wider class of solid celestial objects such as Earth-like planets. One of the remarkable findings of our investigation is that the dipole overtones of nodeless elastic shear vibrations trapped in the finite-depth crust of seismically active neutron star possess properties of Goldstone soft modes. It is shown that obtained spectral equations are consistent with the existence treatment of low-frequency QPOs in the X-ray luminosity of flares SGR 1900+14 and SGR 1806-20 as caused by quake-induced torsional nodeless vibrations (Samuelsson \\& Andersson 2007a; 2007b). What is newly disclosed here is that previously non-identified low-frequency QPOs in data for SGR 1806-20 can be attributed to nodeless dipole torsional and spheroidal vibrations, namely, $\\nu(_0t_1)=18\\,\\mbox{Hz}$ and $\\nu(_0s_1)=26\\,\\mbox{Hz}$." }, "0710/0710.5508_arXiv.txt": { "abstract": "We analyse a sample of 33 extensive air showers (EAS) with estimated primary energies above $2\\cdot 10^{19}$~eV and high-quality muon data recorded by the Yakutsk EAS array. We compare, event-by-event, the observed muon density to that expected from CORSIKA simulations for primary protons and iron, using SIBYLL and EPOS hadronic interaction models. The study suggests the presence of two distinct hadronic components, ``light'' and ``heavy''. Simulations with EPOS are in a good agreement with the expected composition in which the light component corresponds to protons and the heavy component to iron-like nuclei. With SYBILL, simulated muon densities for iron primaries are a factor of $\\sim 1.5$ less than those observed for the heavy component, for the same electromagnetic signal. Assuming two-component proton-iron composition and the EPOS model, the fraction of protons with energies $E>10^{19}$~eV is $0.52 ^{+0.19}_{-0.20}$ at 95\\% confidence level. ", "introduction": " ", "conclusions": "" }, "0710/0710.4162_arXiv.txt": { "abstract": "Estimates of inflationary parameters from the CMB $B$-mode polarization spectrum on the largest scales depend on knowledge of the reionization history, especially at low tensor-to-scalar ratio. Assuming an incorrect reionization history in the analysis of such polarization data can strongly bias the inflationary parameters. One consequence is that the single-field slow-roll consistency relation between the tensor-to-scalar ratio and tensor tilt might be excluded with high significance even if this relation holds in reality. We explain the origin of the bias and present case studies with various tensor amplitudes and noise characteristics. A more model-independent approach can account for uncertainties about reionization, and we show that parametrizing the reionization history by a set of its principal components with respect to $E$-mode polarization removes the bias in inflationary parameter measurement with little degradation in precision. ", "introduction": "\\label{sec:intro} Temperature and polarization power spectra of the cosmic microwave background (CMB) are consistent with predictions of the simplest inflationary models~\\cite{Gut81,AlbSte82,Lin82,Sat81}: a nearly flat geometry, superhorizon correlations as probed by the spectrum of acoustic peaks, and primordial scalar perturbations that are adiabatic, Gaussian, and close to scale-invariant~\\cite{HuWhi96c,SpeZal97,HuSpeWhi97,Peietal03,Speetal07}. One of the key remaining signatures of inflation, tensor perturbations (i.e.\\ gravitational waves)~\\cite{KamKosSte97,SelZal97}, has yet to be detected. Depending on the amplitude of the tensor perturbations, which is not well constrained theoretically, it may be possible to measure the angular power spectrum of the inflationary gravitational waves in the $B$-mode component of the CMB polarization on large scales. Non-detection of the tensor spectrum does not necessarily rule out inflation, but upper limits on $r$ can be used to exclude particular models of inflation and limit its energy scale. Many experiments have been proposed to search for this signal~\\cite{Planck,CMBTaskForce,Oxletal04,LawGaiSei04,Ruhl:2004kv,Mafetal05,Yooetal06,Kogut:2006nd,Macetal07,Polenta07}. Measurement of tensor perturbations in the $B$-mode polarization power spectrum would test models of inflation by constraining inflationary parameters. These parameters include the tensor-to-scalar ratio, $r$, and the tensor spectral index, $n_t$, which are related by a consistency relation under the simplest single-field slow-roll inflationary scenarios. If the tensor spectrum can be detected, precise measurements over a wide range of scales could test the consistency relation. CMB constraints on $r$ and $n_t$ depend on the ability to accurately determine the large-scale power in $B$-modes due to tensor perturbations, independent of the effects of other cosmological parameters. On the largest scales, the tensor $B$-mode spectrum depends not only on inflationary parameters but also on the reionization history of the universe~\\cite{Zal97}. The main impact of reionization on the spectrum is through the total optical depth, $\\tau$. The 3-year \\wmap\\ measurements of $E$-mode polarization determine $\\tau$ to an accuracy of about $30\\%$~\\cite{Pagetal07,Speetal07}, and future CMB experiments should constrain $\\tau$ at the~$5-10\\%$~level~\\cite{Holetal03,KeaMil06,Planck,MorHu07b}. However, the detailed evolution of the reionization history also significantly affects the large-scale polarization spectra. Uncertainty in this history leads to added uncertainty in inflationary parameters. Moreover, incorrect inferences due to an oversimplified treatment of reionization may bias estimates of inflationary parameters. For unbiased estimation of the optical depth from the $E$-mode reionization peak, the solution is to use a complete, principal-component-based description of reionization effects when estimating parameters from CMB polarization data~\\cite{HuHol03,MorHu07b}. In this paper, we extend this approach to tensor $B$-mode polarization and show that it is equally if not more effective in ensuring accurate measurements with little loss in precision. The outline of the paper is as follows. We discuss the effects of reionization and inflationary parameters on the polarization power spectra and the large-scale degeneracy between these parameters in \\S~\\ref{sec:param}. A brief overview of the principal component parametrization of the reionization history follows in \\S~\\ref{sec:pcs}. In \\S~\\ref{sec:mcmc}, we describe our Markov Chain Monte Carlo analysis of simulated polarization data and give the resulting constraints on $\\tau$, $r$, and $n_t$, which we discuss further in \\S~\\ref{sec:discuss}. \\begin{figure} \\centerline{\\psfig{file=f1.eps, width=3.0in}} \\caption{$B$-mode tensor spectra illustrating the degeneracy between $\\tau$ and $n_t$ for large-scale measurements, with angular power spectra plotted in the upper panel and fractional deviations from the base model in the lower panel. For the base model (\\emph{solid}), $r=0.03$, $\\tau=0.1$, and $n_t=-0.00375$. The other two models have $\\{\\tau,n_t\\}=\\{0.12,-0.00375\\}$ (\\emph{long dashed}) and $\\{0.12,0.13\\}$ (\\emph{short dashed}), with a pivot scale of $k_{\\rm pivot}=0.01~{\\rm Mpc}^{-1}$. The reionization history here is assumed to be instantaneous. Cosmic variance of $\\cltens$ for the base model, which excludes the variance from lensing, is shown by the shaded band in the lower panel. } \\label{fig:degen} \\end{figure} ", "conclusions": "\\label{sec:discuss} The value of the optical depth to reionization estimated from the CMB $E$-mode polarization spectrum on large scales can be biased by adopting a model that has insufficient freedom to describe the true reionization history. Likewise, the use of simple reionization models can bias inflationary parameters such as the tensor-to-scalar ratio and tensor tilt that depend on the large-scale amplitude of the $B$-mode spectrum of primordial gravitational waves. In each case, the problem can be solved by using a more general parametrization of the reionization history. We have shown that using a small but complete set of the principal components of the reionization history effectively yields unbiased constraints on both reionization and inflationary parameters. Measurements of $r$ and $n_t$ are only affected by the assumed form of the reionization history if the reionization peak of the tensor $B$-mode spectrum at the very largest scales is needed to precisely constrain the parameters. If, instead, good constraints can be obtained using only the $B$-mode recombination peak at intermediate scales, then assumptions about reionization do not affect tests of the consistency relation between $r$ and $n_t$. They would instead appear as false evidence for running of the tensor tilt in violation of slow-roll expectations. Measurement of the recombination peak however is inhibited by experimental noise and contamination from $E$-mode power converted to $B$-mode power by gravitational lensing, both of which become more important at smaller scales. To study the potential impact of reionization on parameter constraints from $B$-mode polarization, we have employed a Markov Chain Monte Carlo analysis of simulated CMB polarization power spectra and compared results for two descriptions of reionization: a simple, one-parameter, instantaneous reionization model, and a parametrization using principal components of the reionization history with respect to the $E$-mode polarization power spectrum. By varying the properties of the simulated polarization power spectra, including the fiducial tensor-to-scalar ratio and the noise spectrum, we have determined over what range of scales CMB polarization data is most important for constraining inflationary parameters in various scenarios. In particular, the question of whether the large-scale reionization peak of $\\cltens$ or the smaller-scale recombination peak is more important determines the severity of bias in inflationary parameters when reionization is modeled incorrectly. If the tensor-to-scalar ratio is near the current upper limit of $r\\sim 0.3$ and measurements of $B$-mode polarization are limited only by cosmic variance, then the spectrum on scales $20 \\lesssim \\ell \\lesssim 500$ dominates constraints on $r$ and $n_t$ and incorrect assumptions about reionization do not strongly bias the results. If the true tensor-to-scalar ratio is more than a factor of a few smaller than this upper bound, however, then lensing $B$-modes limit the information that can be extracted from the recombination peak of the tensor spectrum alone. Furthermore, all-sky experiments in the foreseeable future are likely to have noise that exceeds the lensing signal, making tests of inflation even more reliant on the reionization peak of the tensor $B$-modes on large scales. In all of these cases, a general parametrization of reionization such as that provided by principal components allows the use of the $B$-mode reionization peak for inflationary parameter constraints without significantly worsening the errors on those parameters. \\vfill" }, "0710/0710.4448_arXiv.txt": { "abstract": "The first high resolution \\textit{Spitzer} IRS 9-37$\\mu$m spectra of 29 Seyfert galaxies (about one quarter) of the 12$\\mu$m Active Galaxy Sample are presented and discussed. The high resolution spectroscopy was obtained with corresponding off-source observations. This allows excellent background subtraction, so that the continuum levels and strengths of weak emission lines are accurately measured. The result is several new combinations of emission line ratios, line/continuum and continuum/continuum ratios that turn out to be effective diagnostics of the strength of the AGN component in the IR emission of these galaxies. The line ratios [NeV]/[NeII], [OIV]/[NeII], already known, but also [NeIII]/[NeII] and [NeV]/[SiII] can all be effectively used to measure the dominance of the AGN. We extend the analysis, already done using the 6.2$\\mu$m PAH emission feature, to the equivalent width of the 11.25$\\mu$m PAH feature, which also anti-correlates with the dominance of the AGN. We measure that the 11.25$\\mu$m PAH feature has a constant ratio with the H$_2$ S(1) irrespective of Seyfert type, approximately 10 to 1. Using the ratio of accurate flux measurements at about 19$\\mu$m with the two spectrometer channels, having aperture areas differing by a factor 4, we measured the source extendness and correlated it with the emission line and PAH feature equivalent widths. The extendness of the source gives another measure of the AGN dominance and correlates both with the EWs of [NeII] and PAH emission. Using the rotational transitions of H$_2$ we were able to estimate temperatures (200-300K) and masses (1-10 $\\times$ 10$^{6}$ M$_{\\sun}$), or significant limits on them, for the warm molecular component in the galaxies observed. Finally we used the ratios of the doublets of [NeV] and of [SIII] to estimate the gas electron density, which appears to be of the order of n$_e$ $\\sim$ 10$^{3-4}$ cm$^{-3}$ for the highly ionized component and a factor 10 lower for the intermediate ionization gas. ", "introduction": "Mid-infrared (mid-IR) spectroscopy provides a powerful tool to investigate the nature and physical processes in active galactic nuclei (AGNs) and in the starburst dominated regions frequently associated to them. Because of the large variety of fine structure lines present in the mid-IR, covering a wide range in ionization/excitation conditions and gas density \\citep[e.g.][]{sm92} mid-IR spectroscopy of the Narrow Line Regions (NLR) in AGNs can add information not available from classical optical spectroscopy, especially when dust extinction is high. Furthermore, the electronic transitions responsible for the infrared emission lines of various elements are less sensitive to uncertainties in temperature than the corresponding optical lines. Moreover the brightest H$_2$ rotational lines, that can be used to quantify the presence and excitation of warm molecular gas, as well as a prominent Polycyclic Aromatic Hydrocarbons \\citep{pule89}, hereafter PAH, feature at 11.25 $\\mu$m are in the mid-IR. The first detailed mid-IR spectroscopic studies of AGNs and Starburst galaxies using multiple ionic transitions of various elements were performed by various authors \\citep[][for a review]{stu02, spi05, ver03, ver05} with the \\textit{Short Wavelength Spectrometer} (SWS) \\citep{deg96} onboard of the \\textit{Infrared Space Observatory} (ISO) \\citep{kes96}. However, the improved sensitivity of the Infrared Spectrometer (IRS) \\citep{hou04} on the \\textit{Spitzer Space Telescope} now enables a detailed investigation of the nature and physical processes in large samples of galactic nuclei from nonstellar (AGN) and stellar (starburst) power sources. Even at low resolution, the IRS spectra of the first few classical AGNs already showed the diversity of the mid-IR spectral features: silicate absorption and emission, PAH emission and strong fine structure lines \\citep{wed05}. \\citet{buc06} examined 51 low resolution IRS spectra of 12$\\mu$m selected Seyfert galaxies \\citep{rms93}, exactly the sample we are considering in our study. They report a few major findings: (1) the sample contains a very wide range of continuum types, with no more than about 3 galaxies being closely similar to one another, however principal component analysis applied to their data suggests that the relative contribution of starburst emission may be the dominant cause of variance in the observed spectra; (2) the starburst component in the sample objects does not contribute more than 40\\% of the total IR flux density; (3) Seyfert 1's have higher ratios of infrared to radio emission \\citep[see also][]{rme96}; (4) the Seyfert 2 galaxies typically show stronger starburst contributions than Seyfert 1's. \\citet{Gor07} found a strong correlation between the [NeV]14.3$\\mu$m and the [NeIII]15.5$\\mu$m lines in Narrow Line Regions (NLR) of AGNs, spanning 4 orders of magnitude in luminosity. This would imply a very narrow range in ionization parameter (-2.8$<$logU$<$-2.5) for simple constant density photoionization models. \\citet{dud07} discussed the ratio of the doublets of [NeV] and [SIII] in a heterogeneous sample of active galaxies. Finally ULIRG galaxies, more than Seyfert galaxies, have been so far the object of systematic spectroscopic studies with Spitzer IRS \\citep{hig06,arm07,des07,far07}. In this article we present the first high resolution IRS spectra of 29 galaxies from the original list of Seyfert galaxies of the \"12$\\mu$m Galaxy Sample\" (12MGS) \\citep{rms93}, an IRAS-selected all-sky survey flux-limited to 0.22 Jy at 12$\\mu$m. The sample selection is briefly described in \\S 2, the observations are described in \\S 3, the results are presented in \\S 4, the diagnostics diagrams using line ratios, equivalent widths and other observed quantities are presented and discussed in \\S 5 and a comparison between Seyfert 1's and Seyfert 2's is given in \\S 6. The conclusions are given in \\S 7. ", "conclusions": "We have identified a family of IRS observables which quantify the proportion of the total IR emission coming from a Seyfert nucleus, all of which are intercorrelated with each other. The ratios of ionic fine structure lines [NeV]/[NeII] and [OIV]/[NeII] were already proposed to measure the importance of the AGN component. We also see that [NeIII]/[NeII] and OIV or [NeV]/[SiII]35$\\mu$m can be used to quantify the AGN dominance. It was also known from ISO that the equivalent width of the 6.2$\\mu$m (and 7.7$\\mu$m) PAH features is inversely related to the AGN dominance; we find that the same holds for the equivalent width of the 11.25$\\mu$m PAH feature. We also discovered two additional IRS observables: the equivalent width of [NeII]12.8$\\mu$m and the extendness of the 19$\\mu$m continuum, which also quantify the dominance of the AGN component compared with the emission from the underlying spiral galaxy. All of these observables are correlated with each other, since they are measuring this same astrophysical quantity, and they are all correlated with the hardness of the far-IR continuum, since a more dominant AGN component is already known to be correlated with hotter dust. There is no clear indication that recent star formation is much more important on average in the Seyfert 2's in our sample, compared with the Seyfert 1's. Although the Seyfert 1's generally tend to have more dominant AGN than the Seyfert 2's, there is a strong overlap between these two classes. The relatively small difference between the averages of these observables for Seyfert 1's and Seyfert 2's indicates that those two AGN categories are not extremely different in the mid-IR range. Thus for example, the AGN in Seyfert 2's are clearly observable in their 10$\\mu$m continuum emission, and in [NeIII], [OIV] and [NeV], to almost the same degree as in Seyfert 1's, which is not consistent with some proposed torus models for Seyfert 1/2 unification. Finally, we do not fully confirm the observational claims of \\citet{dud07} : 75 - 80\\% of our Seyfert 1's and Seyfert 2's show [SIII] or [NeV] densities larger than the low-density limit. In fact, after applying plausible aperture corrections to the [SIII] line ratio, only three Seyfert 2's and one Seyfert 1 have a [SIII] line ratio below the density limit. A similar result is also found for the [NeV] ratio, for which we have two Seyfert 1's (that became four including the upper limits) and 2 Seyfert 2's below the low density limit. These few cases do not in our view require enormous dust extinction values." }, "0710/0710.2617.txt": { "abstract": "We consider the propagation of cosmic rays in turbulent magnetic fields. We use the models of magnetohydrodynamic turbulence that were tested in numerical simulations, in which the turbulence is injected on large scale and cascades to small scales. Our attention is focused on the models of the strong turbulence, but we also briefly discuss the effects that the weak turbulence and the slab Alfv\\'enic perturbations can have. The latter are likely to emerge as a result of instabilities with in the cosmic ray fluid itself, e.g., beaming and gyroresonance instabilities of cosmic rays. To describe the interaction of cosmic rays with magnetic perturbations we develop a non-linear formalism that extends the ordinary Quasi-Linear Theory (QLT) that is routinely used for the purpose. This allows us to avoid the usual problem of 90 degree scattering and enable our computation of the mean free path of cosmic rays. We apply the formalism to the cosmic ray propagation in the galactic halo and in the Warm Ionized medium (WIM). In addition, we address the issue of the transport of cosmic rays perpendicular to the mean magnetic field and show that the issue of cosmic ray subdiffusion (i.e., propagation with retracing the trajectories backwards, which slows down the diffusion) is only important for restricted cases when the ambient turbulence is far from what numerical simulations suggest to us. As a result, this work provides formalism that can be applied for calculating cosmic ray propagation in a wide variety of circumstances. ", "introduction": "The propagation and acceleration of cosmic rays (CRs) is governed by their interactions with magnetic fields. Astrophysical magnetic fields are turbulent and, therefore, the resonant and non-resonant (e.g., transient time damping, or TTD) interaction of cosmic rays with MHD turbulence is the accepted principal mechanism to scatter and isotropize cosmic rays (see Schlickeiser 2002). In addition, efficient scattering is essential for the acceleration of cosmic rays. For instance, scattering of cosmic rays back into the shock is a vital component of the first order Fermi acceleration (see Longair 1997). At the same time, stochastic acceleration by turbulence is entirely based on scattering. It is generally accepted that properties of turbulence are vital for the correct description of CR propagation. Historically, the most widely used model is the model composed of slab perturbations and 2D MHD perturbations (see Bieber, Smith, \\& Matthaeus 1988). The advantage of this empirical model is its simplicity and the ability to account for the propagation of CRs in magnetosphere given a proper partition of the energy between the two types of modes. Numerical simulations (see Cho \\& Vishniac 2000, Maron \\& Goldreich 2001, M\\\"uller \\& Biskamp 2000, Cho, Lazarian \\& Vishniac 2002, Cho \\& Lazarian 2002, 2003, see also book of Biskamp 2003, as well as, Cho, Lazarian \\& Vishniac (2003) and Elmegreen \\& Scalo 2004 for reviews), however, do not show 2D modes, but instead show Alfv\\'enic modes that exhibit scale-dependent anisotropy consistent with predictions in Goldreich \\& Sridhar (1995, henceforth GS95). The approach in the latter work makes productive use of the earlier advances in understanding of MHD turbulence, that can be traced back to Iroshnikov (1963) and Kraichnan (1965) work and the classical work that followed (see Montgometry \\& Turner 1981, Higdon 1984, Montgomery, Brown \\& Matthaeus 1987). A careful analysis shows that there is no big gap between the Reduced MHD and the GS95 model. In fact, it was shown in Lazarian \\& Vishniac (1999) that the numerical results in Matthaeus et al. (1998) are consistent with GS95 predictions. While particular aspects of the GS95 model, e.g., the particular value of the spectral index, are the subject of controversies\\footnote{To address quantitatively these controversies, we need much better numerical resolution. For instance, hydrodynamic turbulence simulations in Kritsuk et al. (2007) showed that only starting with the 1028$^3$ numerical cubes the bottleneck effects stop dominating the measured spectral slope. While the simulations in Kowal \\& Lazarian (2007) show that the bottleneck is less important for their MHD code, the exact value of the spectral slope is still uncertain. At the same time, the particular theoretically-predicted features of turbulence, for instance, the existence of the scale-dependent anisotropy, can be reliably established, while the exact scaling of this dependence, e.g., like $k_{\\|}\\sim k_{\\bot}^{2/3}$ as in GS95 model or $k_{\\|}\\sim k_{\\bot}^{\\alpha*}$, where $\\alpha*<2/3$ (see Beresnyak \\& Lazarian 2006), also require higher resolution simulations.} (see M\\\"uller \\& Biskamp 2000, Boldyrev 2005, 2006, Beresnyak \\& Lazarian 2006, Gogoberidze 2006, Mason et al. 2007), we think that, at present, GS95 model provides a good starting for developing models of CR scattering. as was done in Chandran (2000), Yan \\& Lazarian (2002, 2004, henceforth YL02, Paper I, respectively), Brunetti \\& Lazarian (2007) etc. In particular, the latter three papers used the decomposition of MHD turbulence over Alfv\\'en, slow and fast modes as in Cho \\& Lazarian (2003) and identify the fast modes as the major source of CR scattering in interstellar and intracluster medium. However, while the turbulence injected on large scales may correspond to GS95 model and its extensions to compressible medium (Lithwick \\& Goldreich 2001, Cho \\& Lazarian 2002, 2003), one should not disregard the possibilities of generation of additional perturbations by CR themselves. Indeed, the slab Alfv\\'enic perturbation can be created, e.g., via streaming instability (see Wentzel 1974, Cesarsky 1980) or kinetic gyroresonance instability (see its application for CR transport in Lazarian \\& Beresnyak 2006). These perturbations, that are present for a range of CRs energies (e.g., $\\lesssim 100$GeV for the instabilities above in ISM) owing to non-linear damping arising from ambient turbulence (YL02, Paper I, Farmer \\& Goldreich 2004, Lazarian \\& Beresnyak 2006), should also be incorporated into the comprehensive models of CR propagation and acceleration. At present, the propagation of the CRs is an advanced theory, which makes use both of analytical studies and numerical simulations. However, these advances have been done within the turbulence paradigm which is being changed by the current research in the field. As we discussed above, instead of the empirical 2D+slab model of turbulence, numerical simulations suggest anisotropic Alfv\\'enic modes (an analog of 2D, but not an exact one, as the anisotropy changes with the scale involved) + fast modes or/and slab modes. This calls for important revisions of the CR propagation, which is the subject of the current paper. The perturbations of turbulent magnetic field are usually accounted for by direct numerical scattering simulations (Giacalone \\& Jokipii 1999, Mace et al 2000, Qin at al. 2002) or by quasi-linear theory, QLT (see Jokipii 1966, Schlickeiser 2002). The problem with direct numerical simulations of scattering is that the present-day MHD simulations have rather limited inertial range. At the same time, creating synthetic turbulence data which would correspond to scale-dependent anisotropy in respect to the local magnetic field (which corresponds, e.g., to GS95 model) is challenging and has not been practically realized, as far as we know. While QLT allows easily to treat the CR dynamics in a local magnetic field system of reference, a key assumption in QLT, that the particle's orbit is unperturbed, makes one wonder about the limitations of the approximation. Indeed, while QLT provides simple physical insights into scattering, it is known to have problems. For instance, it fails in treating $90^o$ scattering (see Jones, Birmingham \\& Kaiser 1973, 1978; V\\\"olk 1973, 1975; Owens 1974; Goldstein 1976; Felice \\& Kulsrud 2001) and perpendicular transport (see K\\'ota \\& Jokipii 2000, Matthaeus et al. 2003). Indeed, many attempts have been made to improve the QLT and various non-linear theories have been attempted (see Dupree 1966, V\\\"olk 1973, 1975, Jones, Kaiser \\& Birmingham 1973, Goldstein 1976). Currently we observe a surge of interest in finding way to go beyond QLT. Those include recently developed nonlinear guiding center theory (see Matthaeus et al. 2003), weakly nonlinear theory (Shalchi et al. 2004), second-order quasilinear theory (Shalchi 2005a) (see also Shalchi 2006, Webb et al. 2006, Qin 2007, Le Roux \\& Webb 2007). At the same time, most of the analysis so far has been confined to traditional 2D+slab models of MHD turbulence. Following the reasoning above, we think that it is important to extend the work to the non-linear treatment of CR scattering to models MHD turbulence that are supported by numerical simulations. Propagation of CRs perpendicular to the mean magnetic field is another important problem in which QLT encounters serious difficulties. Compound diffusion, resulting from the convolution of diffusion along the magnetic field line and diffusion of field line perpendicular to mean field direction, has been invoked to discuss transport of cosmic rays in the Milky Way (Getmantsev 1963; Lingenfelter, Ramary \\& Fisk 1971; Allan 1972). The role of compound diffusion in the acceleration of CRs at quasi-perpendicular shocks were investigated by Duffy et al. (1995) and Kirk et al. (1996). Indeed, the idea of CR transport in the direction perpendicular to the mean magnetic field being dominated by the field line random walk (FLRW, Jokipii 1966, Jokipii \\& Parker 1969, Forman et al. 1974) can be easily justified only in a restricted situation where the turbulence perturbations are small and CRs do not scatter backwards to retrace their trajectories. If the latter is not true, the particle motions are subdiffusive, i.e., the squared distance diffused growing as not as $t$ but as $t^{\\alpha}$, $\\alpha<1$, e.g., $\\alpha=1/2$ (K\\'ota \\& Jokipii 2000, Mace et al 2000, Qin at al. 2002, Shalchi 2005b). If true, this could indicate a substantial shift in the paradigm of CR transport, a shift that surely dwarfs a modification of magnetic turbulence model from the 2D+slab to a more simulation-motivated model that we deal here. It was also proposed that with substantial transverse structure, {\\it i.e.}, transverse displacement of field lines, perpendicular diffusion is recovered (Qin et al 2002). Is it the case of the MHD turbulence models we deal with? How realistic is the subdiffusion in the presence of turbulence? The answer for this question apparently depends on the models of turbulence chosen. In this paper we again seek the answer for this question within domain of numerically tested models of MHD turbulence. There are three major thrusts of the paper:\\\\ I. Extend QLT by taking into account magnetic mirroring effect on large scales.\\\\ II. Describe CR propagation in Milky Way (e.g., calculate CR mean free path for different phases of ISM).\\\\ III. Address the problem of perpendicular transport of CR. In what follows, we discuss the cosmic ray transport in incompressible turbulence in \\S2. We shall describe the \\S2.1 dispersion of guiding center of CRs and introduce the broadened resonance function to replace the $\\delta$ function in QLT, following which we shall discuss the scattering in strong and weak incompressible turbulence respectively in \\S2.2 and \\S2.3. Then we shall consider the scattering by fast modes in \\S3 and apply the analysis to ISM and get mean free path for different phases of ISM (\\S4). In \\S5, we shall study the perpendicular transport of cosmic rays on both large and small scales. We shall also discuss the applicability of the subdiffusion. Discussion and summary are provided in \\S6 and \\S7 respectively. ", "conclusions": "The present paper extends our study in Paper I. As in Paper I we mostly deal with the magnetic perturbations that are part of the large scale turbulent cascade, which is consistent with the Big Power Law in the sky observed via radio-scattering and scintillation technique (Armstrong et al. 1995). In both papers we use the description of the MHD turbulence that follows from numerical simulations. In Paper I we have the CR scattering calculated in the selected interstellar environments making use of Quasi-Linear Theory (QLT). Because of the limitations of the QLT, we could not provide calculations of the mean free path in Paper I, which limited the utility of the study. In this paper we extended the non-linear approach suggested in V\\\"olk (1975) to treat the scattering, which allows us to calculate the mean free paths that arise from CR interactions with the fast modes. In doing so, similar to Paper I, we take into account damping of the fast modes in the presence of the field wondering induced by the Alfv\\'enic modes. Our results show that in WIM and halo of our Galaxy, confinement of bulk CRs are mostly due to the compressible modes. We obtain CR mean free paths about a few parsec, consistent with what observations indicate. The major difference with earlier picture is the dependencies of CR transport parameters on the medium properties. The dependence appears as a result of damping of the fast modes. For low energy CRs ($\\lesssim 100$GeV), if dominated by viscous damping, the mean free path of CRs would decrease with energy; with collisionless damping, however, CRs' mean free path stays almost a constant. Field line wandering in general increases the damping of the fast modes and reduces the scattering efficiency of CRs. For higher energy CRs, the influence of damping is limited, and their mean free path increases with energy. The dependencies on the turbulence damping and therefore the phase properties should have various implications from ratio of secondary to primary elements, diffuse Galactic $\\gamma$ ray emission, to the CMB synchrotron foreground. With precise measurements, the understanding of CMB is now constrained by our understanding of the foreground. The variation of CR index over the Galaxy may paralyze the synchrotron templates. Such variations can be addressed on the basis of the more elaborate CR propagation theory. The importance of this study goes beyond the interstellar medium. For instance, Brunetti \\& Lazarian (2007), treated acceleration of CRs for plasma in clusters of galaxies appealing to the fast modes, which is the approach to CRs similar to that in Paper I. We believe that the non-linear treatment may be useful for such cases as well. In addition, stochastic acceleration by the MHD turbulence is a promising mechanism for generating high energy particles during solar flares (see, e.g., Petrosian \\& Liu 2004, and references therein). An application to the acceleration of CRs in solar flares will be given in Yan, Lazarian \\& Petrosian (2007, in preparation). In our treatment we attempted to use the scalings that (a) are consistent with numerical calculations and (b) whose amplitudes we can estimate with a sufficient degree of precision. Therefore our present study does not deal with scattering of CRs by the fast modes on the scales $l>LM_A^2$, $M_A<1$, i.e., on the scales where the Alfv\\'enic turbulence in the weak regime. It was suggested by Chandran (2005) that the weak fast modes at small pitch angles tend to steepen due to the coupling with the Alfv\\'en modes. When the resulting scaling of the fast modes becomes clearer, our approach will be applicable to them. We have not quantitatively dealt in the present paper with the case of the slab Alfv\\'en modes created by instabilities\\footnote{Streaming instability (see Cesarsky 1980) is an example of such instability. However, the instability is suppressed by both ion-neutral damping (Kulsrud \\& Pierce 1969) and the ambient turbulence (YL02, Farmer \\& Goldreich 2004, Paper I, Lazarian \\& Beresnyak 2006). Another example is the gyroresonance instability discussed in the context of CRs in Lazarian \\& Beresnyak (2006).}. The CR scattering by the perturbations created by those modes may dominate over the gyroresonance with the fast modes, especially for CRs of low energies, i.e., whose gyroresonance with the fast modes is inefficient due to the fast modes damping (see estimates in Lazarian \\& Beresnyak 2006). Progress in quantitative description of the non-linear stages of the instabilities that can create slab modes should enable comprehensive models that include both the fast modes and the slab modes. In addition, we addressed the issue of perpendicular diffusion, the issue that we have not dealt with in Paper I. We found, that similar to the case of thermal diffusion discussed in Lazarian (2006), the diffusion of CRs depends on the Alfv\\'enic Mach number $M_A$. We found that the suppression of the perpendicular diffusion compared to the parallel one scales as $M_A^4$ for $M_A<1$. Approaching the issue of subdiffusion, we found that it is negligible for CRs in the Alfv\\'enic turbulence." }, "0710/0710.4354_arXiv.txt": { "abstract": "We present deep WIYN H$\\alpha$ imaging of the dwarf irregular starburst galaxy NGC 1569, together with WIYN SparsePak spatially-resolved optical spectroscopy of the galactic outflow. This leads on from our previous detailed analyses of the state of the ISM in the central regions of this galaxy. Our deep imaging reveals previously undetected ionized filaments in the outer halo. Through combining these results with our spectroscopy we have been able to re-define the spatial extent of the previously catalogued superbubbles, and derive estimates for their expansion velocities, which we find to be in the range 50--100~\\kms. The implied dynamical ages of $\\lesssim$\\,25~Myr are consistent with the recent star- and cluster-formation histories of the galaxy. Detailed decomposition of the multi-component H$\\alpha$ line has shown that within a distinct region $\\sim$$700\\times 500$~pc in size, roughly centred on the bright super star cluster A, the profile is composed of a bright, narrow (FWHM $\\lesssim$ 70~\\kms) feature with an underlying, broad component (FWHM $\\sim$ 150~\\kms). Applying the conclusions found in our previous work regarding the mechanism through which the broad component is produced, we associate the faint, broad emission with the interaction of the hot, fast-flowing winds from the young star clusters with cool clumps of ISM material. This interaction generates turbulent mixing layers on the surface of the clouds and the evaporation and/or ablation of material into the outflow. Under this interpretation, the extent of the broad component region may indicate an important transition point in the outflow, where ordered expansion begins to dominate over turbulent motion. In this context, we present a multi-wavelength discussion of the evolutionary state of the outflow. ", "introduction": "Outflows powered by the collective injection of kinetic energy and momentum from massive stars and supernovae (SNe) can drastically affect the structure and subsequent evolution of galaxies. Thus, a good understanding of the feedback mechanisms between massive stars, star clusters and the ISM is fundamental. In particular, dwarf galaxies are thought to be strongly affected by the effects of feedback since their smaller gravitational potentials mean that supernova-heated gas can escape more easily \\citep{larson74}. Although the ejection of the ISM potentially could have significant consequences on the star-formation rate within these low-mass systems \\citep{dekel86}, more recent work suggests that ejection of hot gas through bubble blow-out may not be as efficient as first thought \\citep{deyoung94, martin98}. It is therefore important to study such systems to understand how gas is removed and what effects this has in the evolution of the galaxy. NGC 1569 (UGC 3056, Arp 210, VII Zw 16, IRAS 4260+6444) is a nearby \\citep[2.2~Mpc;][]{israel88}, low metalicity \\citep*[0.25~\\Zsun;][]{devost97, kobulnicky97} dwarf irregular galaxy that has recently undergone a period of starburst activity. The most recent burst is thought to have peaked between $\\sim$10--100~Myr ago with an average star-formation rate of $\\sim$0.5~\\Msun~yr$^{-1}$ \\citep{greggio98}. At some point near the end of this event, the two well-known super-star clusters (SSCs) A and B were formed (\\citealt{arp85}; \\citealt*{oconnell94}; \\citealt{demarchi97}), and together with the slightly older cluster 30 \\citep{hunter00, origlia01}, appear to dominate the energetics of the central regions. H\\one\\ observations of NGC 1569 \\citep{stil98, stil02, muhle05} clearly show morphological and kinematical signatures caused by the starburst. Firstly, the H\\one\\ kinematics are ``strongly disrupted'' within the central 900~pc, with little or no evidence for the disc rotation seen at lager radii. Furthermore, large seemingly tidal structures are seen, including a so-called bridge connecting the galaxy to a low-mass H\\one\\ cloud \\citep[the companion;][]{stil98}, a large H\\one\\ arm extending to the west of the disc, and a very faint filamentary H\\one\\ stream wrapping around the south of the disc \\citep{muhle05}. A `hot spot' (a region of velocity crowding) on the western edge of the disk was found by \\citet{muhle05}, who interpreted it as the impact location of infalling gas from the companion. This provides a compelling explanation of how the starburst event was triggered. H$\\alpha$ images of this galaxy show an equally chaotic, complex structure to the warm ionized component, exhibiting many filaments, bubbles and loops. Many of these were identified by \\citet*{hunter93} from deep H$\\alpha$ imaging. Later kinematical studies found that these filaments form part of a cellular outflow structure, comprising large-scale expanding superbubbles on the northern and southern sides of the disc (\\citealt*{tomita94}; \\citealt{heckman95, martin98}). By analysing spectra from two long-slits placed parallel to the major and minor axes, \\citet{heckman95} found evidence of shocks in the outer regions of the ionized halo. The ratios of [O\\one]/H$\\alpha$, [S\\two]/H$\\alpha$ and [N\\two]/H$\\alpha$ \\citep[used often to diagnose and trace the conditions within ionized gas;][]{veilleux87, dopita00} were all found to be high in the outer filaments, indicating either an increase in the importance of shocks, or that photoionization becomes less important with distance as the ionizing radiation from the central starburst becomes diluted. The existence of shocked gas is supported by high-resolution X-ray observations of NGC 1569. \\citet*{martin02} examined the X-ray properties of the outflow with \\textit{Chandra}, and found significant soft (0.3--0.7~keV), diffuse emission coincident with the H$\\alpha$ morphology. Although they found the X-ray colour variations to be inconsistent with a free-streaming wind, they concluded that the X-ray emission probably originates in the halo shock generated by the outflowing gas, possibly from the mixing layers between the shock and the bubble interior. In order to properly characterise the structure of the outflow, it is essential to study its kinematics. From optical long-slit spectroscopy, \\citet{martin98} found the outflow to be composed of numerous superbubbles. In general, she finds the redshifted component of the split-line profile to be stronger in the south and weaker in the north, suggesting an inclined outflow aspect. This is consistent with more recent X-ray absorption measurements \\citep{martin02} and H\\one\\ observations \\citep[from which an inclination angle of $\\sim$60$^{\\circ}$ was derived][]{stil02}. Although it is unclear whether the two SSCs, A and B, alone are sufficiently powerful to provide enough mechanical energy and ionizing radiation to drive the whole outflow and power the galaxy's diffuse ionized medium \\citep{martin97, martin02}, what is clear from the H$\\alpha$ morphology is that energy is being injected throughout the central starburst zone of the disc from multiple sources. A detailed look at the spectral line profiles, however, shows that near SSC A, ``distinctly non-Gaussian'' H$\\alpha$ emission can be found, exhibiting weak but very broad wings \\citep{heckman95}. Broad emission line wings have been detected in other nearby starburst galaxies (e.g.~\\citealt{izotov96, homeier99, marlowe95, mendez97}; \\citealt*{sidoli06}; \\citealt{westm07c}). Due to mismatches in spectral and spatial resolution and in the specific environments observed, the nature of the energy source for these broad lines has been contested, and has resulted in the proposal of number of possible explanations. However, a detailed IFU (integral field unit) study of the ionized ISM conditions in four regions within the 200~pc region surrounding SSC A by \\citet[][hereafter \\citetalias{westm07a}]{westm07a} and \\citet[][hereafter \\citetalias{westm07b}]{westm07b} have shed a considerable amount of light on this problem. By accurately decomposing the emission line profiles across each of the IFU fields, we found the line shape to be, in general, composed of a bright narrow feature (intrinsic FWHM $\\sim$ 50~\\kms) superimposed on a fainter broad component (FWHM $\\sim$ 200--400~\\kms). By mapping out their individual properties, we identified a number of correlations between the line components that allowed us to investigate in detail the state of the ionized ISM. We concluded that the broad underlying component is most likely produced in a turbulent mixing layer \\citep{slavin93, binette99} on the surface of cool gas knots, set up by the impact of the fast-flowing cluster winds \\citep{pittard05}. Our analysis revealed a very complex environment with many overlapping and superimposed components, but surprisingly no evidence for organised bulk motions. We concluded that the four regions sampled are all located well within the wind energy injection zone \\citep{shopbell98} at the very roots of the outflow, and that the collimation processes required to transform the turbulent motions into an organised net outflow forming the large-scale superbubbles must occur between 100--200~pc from the central star clusters. With this in mind, we have obtained new deep H$\\alpha$ imaging of NGC 1569, together with spatially-resolved spectra of the outer halo, to investigate in detail the morphology and kinematics (including the line profile shapes) of the warm ionized component of the outflow at these large radii. In Section~\\ref{sect:data} we present the observations and in Section~\\ref{sect:maps} we map out the properties of the line components and discuss the results of the spectroscopy. Since a number of SparsePak fibres are coincident with or lie adjacent to the Gemini GMOS/IFU fields presented in Papers I and II, we compare the results obtained with the two instruments in Section~\\ref{sect:comparison}. In Section~\\ref{sect:disc} we discuss the state of the outflow, including the conditions in the inner and outer halo, and we summarise our findings and conclusions in Section~\\ref{sect:concs}. ", "conclusions": "\\label{sect:concs} We have presented WIYN MiniMo deep H$\\alpha$ imaging of NGC 1569 covering a field-of-view of $9.5\\times 10.7$~arcmins. The depth and large dynamic range of the observations have enabled the identification of previously undetected faint ionized filaments in the halo and the study of the gas morphology right down into the bright, central regions of the disc. We have also presented WIYN SparsePak ``formatted field unit'' observations covering the outer galactic wind flow of NGC 1569 in four pointings with integration times of 4--4.5 hours per field. The large diameter of the SparsePak fibres makes this instrument ideal for probing the faint ionized gas found in the halos of galaxies. This light-collecting power allowed us to choose a high-resolution spectrograph set-up, enabling us to characterise the line profile shapes of the important nebular diagnostic lines of H$\\alpha$ and [S\\two] to an accuracy limited only by the S/N achieved. We now summarise our main findings. \\begin{itemize} \\item We find H$\\alpha$ emission out to radii of $\\sim$1.5~kpc from the disc, and detect emission in almost every SparsePak fibre over the combined FoV. The presence of such an extensive system of ionized filaments results from the ongoing starburst that is supplying both mechanical energy to eject material and the Lyman continuum luminosity to keep it photoionized. \\item Through detailed Gaussian line fitting, we find that within a distinct region $\\sim$$700\\times 500$~pc in size, roughly centred on the location of SSC A, the nebular emission line profiles are composed of a bright, narrow (FWHM $\\lesssim$ 70~\\kms) component with an underlying, broad component (FWHM $\\sim$ 150~\\kms). At larger radii, we find two narrow components to the H$\\alpha$ line, each representing one half of a split-line profile. \\item By comparing our results to observations of the central regions directly surrounding SSC A \\citepalias{westm07a, westm07b}, we conclude that the physical mechanisms that give rise to the underlying broad emission seen within this zone must be the same as within the regions directly surrounding SSC A sampled by Gemini GMOS/IFU observations. The broad emission is most likely to result from turbulent mixing layers on the surface of cool gas clumps set up by the impact of the hot, fast-flowing cluster winds, and from evaporation and/or ablation of material from the clumps. \\item The extent of this broad component region, coincident with the point at which we start observing signatures of large-scale bubble expansion, may indicate a transition point where ordered expansion begins to dominate over turbulent motion. Further observations are needed to investigate this in more detail. \\item By combining our deep H$\\alpha$ imaging and spectroscopy, we redefine the spatial extents of superbubbles A and B, and confirm their published \\citep{martin98} expansion velocities (90 and 85~\\kms, respectively). We estimate the dynamical ages of these bubbles to be $\\sim$10--15~Myr. Contrary to what has been previously suggested \\citep{martin98}, we find no kinematic or morphological evidence to suggest that either of these two superbubbles have ruptured and are venting their interiors into the galactic halo. \\item Our data indicate that the halo of NGC 1569 contains only 4 superbubbles. Following the terminology introduced by \\citet{martin98}, we propose that her superbubble F should should encompass the whole south-eastern bubble complex, where the velocity ellipses identified by \\citet{martin98} and used to define shells G and F, are simply one level of a `hierarchy of structure'. Furthermore, we propose that superbubble E should encompass what were previously referred to as shells D and E and the large north-eastern X-ray spur \\citep{martin02}. \\item We derive new measurements of the expansion velocity, $v_{\\rm exp}$ (calculated from the difference in the radial velocities between the two split-line components), for the superbubble complexes E and F of $v_{\\rm exp}\\sim50$~\\kms{} and $\\lesssim$100~\\kms, respectively. Assuming a diameter of $\\sim$1.3~kpc for these two structures implies dynamical ages of $\\lesssim$25~Myr. \\item The derived ages of the supershells are consistent with the recent cluster formation history of NGC 1569 \\citep{anders04}, implying that each shell is associated with a specific star-forming event, such as a young massive star cluster that can provide a large mechanical luminosity from its many type II supernovae. \\item The consistent reversal of strengths between the blue and red components in the northern and southern outflows provides evidence of preferred outflow directions approximately perpendicular to the inclined and flattened H\\one\\ disc \\citep{stil02}. \\item In addition to characterising the H$\\alpha$ line profile, we have also measured [S\\two] derived electron densities and [S\\two]$\\lambda$6717+$\\lambda$6731/H$\\alpha$ line ratios. We find that much of the ionized gas in the broad component region and in the outer-wind regions is at or below the low density limit, as is expected for a rarefied outflow. We find the highest [S\\two]/H$\\alpha$ ratios are associated with the faintest H$\\alpha$ fluxes and the largest galactocentric distances. log([S\\two]/H$\\alpha$) ratios $>$0 are found in the superbubbles E, B and A, indicating that in these regions the gas emission may be significantly shock-excited. \\end{itemize} In summary, the outflow in NGC~1569 appears to consist of several superbubbles in various phases of development, as noted by \\citet{martin98}. This situation is further complicated by the disturbed state of the H\\one\\ that includes features well out of the main plane of the galaxy (Fig.~\\ref{fig:xray_HI}b). Our data confirm this model and thus indicate that the evolution of the outflow will be determined by the development of the superbubbles. In superbubbles A, B, and F ionized gas arcs seen in H$\\alpha$ and the X-ray morphology suggest that much of the hot gas still is confined, albeit moving at velocities that are comparable to those needed to escape. Thus the NGC~1569 outflow does not currently appear to be in the form of an approximately steady state galactic wind, even though it may eventually lead to mass loss from the system. It therefore differs from the well known M82 outflow, whose outer regions can be modelled by a supersonic galactic wind \\citep[e.g.][]{suchkov94, shopbell98, zirakashvili06}." }, "0710/0710.2003_arXiv.txt": { "abstract": "The apparent alignment of the cosmic microwave background multipoles on large scales challenges the standard cosmological model. Scalar field inflation is isotropic and cannot account for the observed alignment. We explore the imprints, a non-standard spinor driven inflation would leave on the cosmic microwave background anisotropies. We show it is natural to expect an anisotropic inflationary expansion of the Universe which has the effect of suppressing the low multipole amplitude of the primordial power spectrum, while at the same time to provide the usual inflationary features. ", "introduction": " ", "conclusions": "" }, "0710/0710.0837_arXiv.txt": { "abstract": "\\citet{chang06} reported millisecond duration dips in the X-ray intensity of Sco X-1 and attributed them to occultations of the source by small trans-Neptunian objects (TNOs). We have found multiple lines of evidence that these dips are not astronomical in origin, but rather the result of high-energy charged particle events in the {\\it RXTE} PCA detectors. Our analysis of the {\\it RXTE} data indicates that at most 10\\% of the observed dips in Sco X-1 could be due to occultations by TNOs, and, furthermore, we find no positive or supporting evidence for any of them being due to TNOs. We therefore believe that it is a mistake to conclude that any TNOs have been detected via occultation of Sco X-1. ", "introduction": "\\label{sec:intro} \\citet{chang06} found statistically significant 1-2 millisecond duration dips in the count rate during X-ray observations of the bright X-ray source Sco X-1 carried out with the Proportional Counter Array (PCA) on the {\\it Rossi X-ray Timing Explorer} ({\\it RXTE}) and attributed them to occultations of the source by small objects orbiting the Sun beyond the orbit of Neptune, i.e., trans-Neptunian objects (TNOs). In all, \\citet{chang06} found some 58 dips in approximately 322 ks of Sco X-1 observations. Given that the {\\it RXTE} spacecraft moves through the diffraction-widened shadows of any TNOs at a velocity of $\\sim30$ km s$^{-1}$, dips of $\\sim$2 ms duration should correspond to a TNO size of $\\sim$60 m. If the identification of these dips with occultations by TNOs is correct, the dips would provide extremely valuable information on the number and distribution of solar system objects of $\\sim20$-100 m in size. We have found evidence that these dips are produced by electronic dead-time as a result of high-energy charged particle events in the RXTE PCA detectors. Preliminary reports of our results were given by \\citet{tajatel, tajast}; herein we give a more detailed and complete report. ", "conclusions": "\\label{sec:disc} The observed dips have widths and depths that are approximately what one might expect to be produced by occultations by TNOs, even though much wider dips would be detectable in principle (given appropriate depths). Thus we are obligated to seriously consider the hypothesis that some or all of the observed dips are the product of TNO occultations. However, close examination of the {\\em RXTE} PCA data reveals six signatures that independently indicate that few and possibly none of the observed dips are due to occultations by TNOs. The signatures are (1) the numbers of SM1-other events during the dips; (2) the numbers of VLE events during the dips; (3) the absence of the expected diffraction sidelobes; (4) the temporal asymmetry of the dips; (5) the almost total lack of dips longer than $\\sim$1 ms; and (6) the lack of correlation between dip duration and depth. We discuss each of these in turn. (1) {\\bf SM1-Other Events:} Fig.~\\ref{fig:xraylc} shows that there is, on average for 201 dips, a large excess of SM1-other events at or near the times of the dips. On average, individual dips should show an excess of SM1-other events at the $2.7\\, \\sigma$ level. In Fig.~\\ref{fig:otherhist} we show a histogram of the number of SM1-other events in the 1/8-s time bins corresponding to the dips expressed in standard deviations from the mean. The mean value obtained from the histogram is $\\sim2.7\\, \\sigma$, as expected. If one makes the reasonable assumption that the numbers of SM1-other events should not be affected by true occultations (other than by negligible increases due to reductions in the electronic dead time), then one may estimate the maximum fraction, $f_{\\rm occ}$, of the observed dips that represent genuine occultation events that is consistent with this distribution of SM1-other events. We constructed the following simple function with which to fit the histogram, and thereby constrain $f_{\\rm occ}$: \\begin{equation} {\\rm probability} = \\frac{1}{\\sqrt{2\\pi}}f_{\\rm occ} e^{-C^2/2} + \\frac{1}{\\sqrt{2\\pi \\sigma_{cr}^2}}(1-f_{\\rm occ})e^{-(C- \\overline{C}_{cr})^2/2\\sigma_{cr}^2} \\end{equation} where $C$ is the number of excess SM1-other counts (in units of standard deviations of the counts per bin in each PCA light curve), and $\\overline{C}_{cr}$ is the mean of $C$ for those dips which are not the results of occultation events and which we take to be $\\approx2.7/(1-f_{\\rm occ})$. The distribution of the numbers of excess SM1-other counts is wider than what would be expected from a Poisson distribution with a mean equal to the slightly increased (on average) number of SM1-other events per bin; the width of this component of the fitting function is adjusted by means of the parameter $\\sigma_{cr}$. Fits of the function to the histogram in Fig.~\\ref{fig:otherhist} were carried out with $\\chi^2$ fits using both Gaussian and Cash (1979) statistics. If we neglect the tail of the distribution at high numbers of excess SM1-other events, i.e., at $>9\\sigma$, we obtain formally acceptable fits with values of $f_{\\rm occ}$ in the range 0.0 to 0.12 and values of $\\sigma_{cr}$ in the range 1.85 to 2.35 (based on Gaussian statistics; the limits represent the formal joint 95\\% confidence range). Using Cash statistics, we obtain formally acceptable fits with values of $f_{\\rm occ}$ in the range 0.0 to 0.11 and values of $\\sigma_{cr}$ in the range 1.65 to 2.15 (95\\% confidence). These results indicate that fewer than 11\\% of the 203 dips might be the product of TNO occultations. \\begin{figure*} \\epsscale{0.85} \\plotone{f7.eps} \\caption{Histogram of the numbers of excess SM1-other events in 1/8-s time bins corresponding to the times of 201 of the 203 dips. The number of excess SM1-other events is given in terms of the square root of the mean number of SM1-other events per bin away from the time of the dip. The solid smooth curve represents the best fit with $f_{\\rm occ} = 0$, $\\sigma_{cr} = 2.1$ and the dashed curve represents a formally {\\em un}acceptable fit with $f_{\\rm occ} = 0.24$, $\\sigma_{cr} = 1.8$ (see text). } \\label{fig:otherhist} \\end{figure*} (2) {\\bf VLE events:} Figure~\\ref{fig:xraylc} also shows that there is an excess of VLE events around the times of the dips. The difference between the background rate and that in the 1/8-s bin containing the dips is very close to 4 (actually $3.9 \\pm 0.5$) extra VLE events per dip. The peak in Fig.~\\ref{fig:xraylc} is significant at the $7 \\sigma$ level. If there is precisely one VLE per operating PCU for each non-TNO dip, then we would expect on average an excess of 4.67 VLEs per non-TNO dip. If only the non-TNO dips contribute to the excess VLE events then there is an upper limit to $f_{\\rm occ}$ that is consistent with the observations. If we further allow that the statistical mean excess number of VLE events per dip may have been as small as 2.9, then a simple calculation gives the limit $f_{\\rm occ} \\lesssim 0.38$ (95\\% confidence). This limit is weaker than for the SM1-other events, and, furthermore, is compromised by the possibility that more than one VLE event could be produced in a operating PCU in a single cosmic-ray induced dip. (3) {\\bf Lack of diffraction sidelobes:} In Fig.~\\ref{simprof} we showed an average profile of simulated dips that had been inserted into actual PCA data. It should be compared to averages of the actual measured dip profiles in Figs.~\\ref{fig_superpos_202} and \\ref{fig_superpos_109}. The average model dip profile shows a clear bump of $\\sim8$\\% amplitude on either side of the dip due to diffraction, whereas the averages of the actual profiles show no significant evidence for diffraction sidelobes. Thus, we conclude that the fraction $f_{\\rm occ}$ of legitimate TNO occultations can be no larger than $\\sim$30\\%, otherwise diffraction sidelobes likely would have been detected. Again, while this is a clear strike against the dips being due to TNOs, the limiting statistically significant constraint that can be set due to the lack of diffraction sidelobes is not as significant as for the SM1-other events. (4) {\\bf Asymmetry:} A comparison of the simulated with the actual dip profiles (as in [3]) above, clearly shows a marked asymmetry for the real dip events. This is physically implausible if the dips are the product of occultation events and therefore testifies against a TNO origin for most of the dip events. We estimate that the statistical significance of the asymmetry is $\\sim 6\\,\\sigma$. Unfortunately, there is no direct way to use this information to constrain the fraction of legitimate TNO occultations. The problem is that we do not know, a priori, how large the asymmetry is, on average, for non-TNO dips. Therefore, we can not tell how `diluted' the non-TNO events are by potentially real ones. Nonethless, this marked asymmetry is another solid indication that few of the dips are the product of TNO occultations. (5) {\\bf Lack of dips longer than $\\sim$1 ms:} From Fig.~\\ref{fig_realwd} we can see that all of the dips, except for a single event, have RMS widths $\\sigma < 1.1$ ms. In Section~\\ref{sec:data} we described a computer simulation of the production, detection, and analysis of dips caused by TNO occultations. For a relative speed between the {\\it RXTE} satellite and the shadows of the putative TNOs of $v_{\\rm rel} = 25$ km s$^{-1}$ we find that the fraction of recovered simulated dips with $\\sigma > 1.1$ ms is $\\sim$27\\%. For $v_{\\rm rel} = 35$ km s$^{-1}$, $9$\\% of the dips have $\\sigma > 1.1$ ms. We estimate that the average relative velocity between {\\it RXTE} and the shadows of any TNOs was not higher than $v_{\\rm rel} \\sim 30$ km s$^{-1}$. For this speed, $16$\\% of the dips are characterized by $\\sigma > 1.1$ ms. Therefore, if {\\em all} of the dips are the result of TNO occultations the number of longer-duration dips should be $\\sim$30, whereas the observed number is actually 1. On the other hand, if only 15\\% of the dips are due to TNO occultations, then we would expect only $\\sim 5$ dips with $\\sigma > 1.1$ ms. This expected number is marginally statistically consistent, i.e., at $\\sim$5\\% confidence, with the detection of one dip with $\\sigma > 1.1$ ms. Therefore, we conclude that the lack of longer dips allows an upper limit of 15\\% to be set on the fraction, $f_{\\rm occ}$, of potentially real TNO occultations. (6) {\\bf Lack of correlation between width and depth:} If the dips were due to TNO occultations of Sco X-1, we would expect a strong correlation between the widths of the dips and their depths. This results from the fact that diffraction produces shallow occultations for the smaller size occulters, while it produces deeper more geometric-shadowing-like occultations for the larger occulters. As can be seen from the distribution of dip widths vs. depths in Fig.~\\ref{fig_realwd}, there is no such correlation, with almost all of the dips confined to a narrow range of widths (between 0.4 and 0.8 ms) and depths that range all the way from 45\\% to nearly 100\\%. Thus, the fact that the dips we detect include a significant number, i.e., $\\sim$20\\%, that are both narrow ($\\sigma < 0.7$ ms) and deep (minimum normalized count rate below 0.2) whereas only $\\sim2$\\% of the `detected' simulated dips (for $v_{\\rm rel} \\sim 30$ km s$^{-1}$) are this narrow and deep, indicates that $\\lesssim10$\\% of the dips might be due to TNO occultations. Given the effects of statistical fluctuations on the observed number of narrow deep dips and the fact that the simulation is based upon somewhat uncertain parameters, it is more reasonable to use these numbers to set an upper limit of $\\sim20$\\%. Summarizing the results from approaches (1) through (6) above, we find limits on the fraction of valid TNOs to be $f_{\\rm occ} <11\\%$, $< 38\\%$, $< 30\\%$, $< Q\\%$, $<15\\%$, $<20\\%$, respectively, where ``Q'' denotes that a formal limit could not be set, but the approach provides an important independent indication that the dips are, for the most part, not the result of TNO occultations. We believe that the combined upper limit on $f_{\\rm occ}$ due to the joint application of all six approaches is simply the minimum value achieved by the most sensitive of these, i.e., the constraints cannot be combined. The reason, in short, is that the effects we explore serve only to statistically limit the number of events which could be due to TNOs rather than to identify specific qualifying events. Therefore, our final limit is simply $f_{\\rm occ} \\lesssim 10\\%$. One might argue, as did \\citet{chang07}, that since $\\sim10$\\% of the observed dips cannot be formally eliminated as being due to TNOs, they serve as viable potential candidates for TNO detections. However, we argue that if 90\\% of the dips can be securely eliminated as TNO occultations, and there are six different and independent indicators that point in the direction of a common cause due to cosmic ray interactions in the detector, then it is most plausible that {\\em all} of the dips have this common origin. While our results cast serious doubt on whether any true occultation events have been detected, one cannot yet conclude with a high degree of confidence that no such events have been detected. Further investigations of the dip phenomenon and its possible causes would be of interest. We are working to obtain a new measurement of, or upper limit on, the rate of occurrence of occultations of Sco X-1 by analyzing the data that are being obtained in a new series of {\\it RXTE} observations of Sco X-1 with high-time-resolution information on VLE events." }, "0710/0710.3252.txt": { "abstract": "% The role of optical Fe\\,{\\sc iii} absorption lines in B-type stars as iron abundance diagnostics is considered. To date, ultraviolet Fe lines have been widely used in B-type stars, although line blending can severely hinder their diagnostic power. Using optical spectra, covering a wavelength range $\\sim$ 3560 -- 9200 \\AA, a sample of Galactic B-type main-sequence and supergiant stars of spectral types B0.5 to B7 are investigated. A comparison of the observed Fe\\,{\\sc iii} spectra of supergiants, and those predicted from the model atmosphere codes {\\sc tlusty} (plane-parallel, non-LTE), with spectra generated using {\\sc synspec} (LTE), and {\\sc cmfgen} (spherical, non-LTE), reveal that non-LTE effects appear small. In addition, a sample of main-sequence and supergiant objects, observed with FEROS, reveal LTE abundance estimates consistent with the Galactic environment and previous optical studies. Based on the present study, we list a number of Fe\\,{\\sc iii} transitions which we recommend for estimating the iron abundance from early B-type stellar spectra. ", "introduction": "%----------------------- Iron lines dominate the spectra of many astrophysical objects, such as novae \\citep{{mck97},{hat07}}, photoionized H II regions \\citep{{rub97},{rod02},{est02}} and active galactic nuclei \\citep*{{sig03},{sig04},{zha06}}. The atomic processes for iron and other iron-group ions have been the subject of numerous investigations, for example the IRON Project \\citep{hum93} which considers applications in astrophysical and laboratory plasmas \\citep{pra96a}. Absorption lines of iron provide important metallicity diagnostics for both stars and galaxies, and also play a key role in investigating star formation histories through element ratios, such as [$\\alpha$/Fe] \\citep{gil98}. However, there is a lack of robust Fe abundance determinations in external galaxies, e.g. the Magellanic Clouds (see for example, \\citealt*{{rol02},{tru02},{tru07}} and \\citealt{mok07}), due to the complex Grotian diagrams for Fe, the low metallicity environment of the Magellanic Clouds and the reliability of the currently available atomic data. Early B-type stars are important for studying the chemical composition of our own and other galaxies \\citep{{kil92},{duf98}}. In the optical spectra of B-type stars, iron lines due to a number of ionization stages are observed (see for example, \\citealt*{{gie92},{len93},{sma97},{mor06}}). Absorption features arising from Fe\\,{\\sc iii} are primarily detected \\citep{har70}, with Fe\\,{\\sc ii} also found in later B-type stars \\citep{pin93}. Fe\\,{\\sc iv} lines are not expected in the optical spectra of B-type stars due to their intrinsic weakness in this temperature range. The optical Fe\\,{\\sc iii} line spectrum has not been widely employed to determine abundances, as it has often been believed to be too weak to provide reliable measurements \\citep{ken94a}. However, it has been used for chemical composition studies of several bright, narrow-lined main-sequence B-type stars, such as $\\zeta$\\,Cas, $\\gamma$\\,Peg, $\\iota$\\,Her, $\\tau$\\,Sco and $\\lambda$\\,Lep \\citep{{sni69},{pet70},{har70},{pet76},{pet85}}. These Galactic objects have sufficiently high Fe content, coupled with narrow metal absorption lines (due to their low projected rotational velocities), so that, even with relatively poor quality optical spectra, Fe\\,{\\sc iii} features can be detected. However, the quality of the available spectra did not allow the study of Fe\\,{\\sc iii} lines in objects of lower metallicity. Therefore, more recently the optical wavelength range has been largely overlooked in favour of the ultraviolet domain \\citep{{swi76},{pet85},{dix98}}, and in particular the very rich spectral region around 1900 \\AA\\ \\citep*{{tho74},{heb83},{ken94a},{gri96},{moe98}}. On the other hand, due to the high density of absorption features and resultant blending, continuum placement in the ultraviolet is difficult, and hence significant errors may be present in the derived abundances \\citep{moe98}. Due to instrumental advances in more recent years, it has become routine to obtain high resolution and signal-to-noise (S/N) spectra, and the use of the optical Fe\\,{\\sc iii} lines as a diagnostic has been revisited (e.g. \\citealt{{cun94},{kil94}}). A number of main-sequence objects have subsequently been re-examined, for example $\\gamma$\\,Peg and $\\iota$\\,Her \\citep*{{pin93},{zon98}}, producing results consistent with the earlier optical analyses. In addition, the higher S/N ratios of the available spectra have made it possible to detect the weak optical Fe\\,{\\sc iii} lines in a number of lower metallicity objects, such as AV\\,304 in the Small Magellanic Cloud \\citep{rol03}, and the globular cluster post-AGB stars ZNG-1 in M\\,10 \\citep{moo04}, Barnard\\,29 in M\\,13 and ROA\\,5701 in $\\omega$\\,Cen \\citep{tho07}. A number of B-type stars have been studied using both ultraviolet and optical spectra, but the corresponding abundance estimates are generally in poor agreement. Estimates from the ultraviolet spectra are consistently lower than those from the optical, for example in the post-AGB stars Barnard\\,29, ROA\\,5701 \\citep{tho07}, BD\\,+33${^\\circ}$2642 \\citep*{nap94} and HD\\,177566 \\citep{{ken94a},{ken94b}}. Young Galactic objects including $\\gamma$\\,Peg \\citep{moe98} and $\\iota$\\,Her \\citep{gri96}, plus stars in known metallicity environments such as the Magellanic Clouds \\citep{duf07}, also display similar discrepancies. Here, high resolution spectra for a number of narrow-lined Galactic B-type main-sequence and supergiant stars, covering a range of spectral types, are analysed using LTE and non-LTE model atmosphere techniques. Details on the observations, models and Fe\\,{\\sc iii} line selection can be found in Sections \\ref{sec_obs} and \\ref{sec_data}. An LTE approximation is considered, and the reliability of such an assumption, along with an assessment of the individual Fe\\,{\\sc iii} lines, is discussed in Section \\ref{sec_discuss}. %----------------------- ", "conclusions": "\\label{sec_conc} %----------------------- Due to their intrinsic weakness, Fe\\,{\\sc iii} absorption lines have not been widely considered for use in chemical composition studies. Instead, the very rich ultraviolet spectral region has been favored. However, the results found in this paper suggest that the optical region can provide consistent results. The stars here display abundance estimates that agree with the Galactic metallicity, and are consistent with previous studies using optical spectra, where available. By contrast, previous determinations from ultraviolet spectra have followed the trends observed in other studies (for example \\citealt{{tho07},{duf07}}), providing lower abundance estimates, in these cases by approximately 0.5 dex. Although our study has concentrated on B-type stars found in the Milky Way, the optical Fe\\,{\\sc iii} lines examined here can be applied to studies of B-type stars in other galaxies, provided suitable S/N spectra are employed. For example, \\citet{tru02} analysed a sample of B-type supergiants in M\\,31, obtaining an Fe abundance from the Fe\\,{\\sc iii} line at 4419 \\AA. \\citet{rol02} analysed a sample of OB-type main-sequence stars from the LMC, finding an Fe abundance for one object (PS\\,34-16), while \\citet{rol03} investigated a B-type dwarf from the SMC (AV\\,304), finding agreement with other giant stars. More recently, \\citet{tru07} used similar methods to those detailed here and obtained Fe abundances for B-type stars in the Galaxy, LMC and SMC, using the two Fe\\,{\\sc iii} lines at 4419 and 4430 \\AA. The results found were consistent with the present accepted metallicities of these systems. These studies indicate that the optical Fe\\,{\\sc iii} lines can provide reliable abundance indicators in different galaxies. Our comparison of stars analysed using the model atmosphere codes {\\sc cmfgen} and {\\sc tlusty} generally shows little difference in the abundance estimates, indicating that the different physical assumptions, in particular non-LTE effects, are small for this species. Therefore, the results suggest that the optical Fe\\,{\\sc iii} absorption line spectrum may be used with confidence in chemical composition studies, and an LTE analysis provides reliable results. The atomic data of \\citet{nah96}, employed both in this paper and by \\citet{cro06}, appear to provide appropriate abundances, although there are some features, such as those at 4005 and 4273 \\AA, whose log\\,{\\it gf} values may be incorrect. Comparing the values in Table \\ref{tab_atdata} shows that, for some features, e.g. the 4166.88, 4419.60 and 5272.90 \\AA\\ lines, there are large differences between the atomic data from the Kurucz database and \\citet{nah96}. Further work is required to refine the atomic data for this species. \\begin{table} \\caption{Recommended Fe\\,{\\sc iii} lines for use as abundance diagnostics.} \\label{tab_touse} \\begin{tabular}{@{}lcccc} \\hline Line\t& Spectral Type\t\t&Line\t& Spectral Type\t\\\\ (\\AA)\t& Range\t\t\t&(\\AA)\t& Range\t\t\\\\ \\hline 4419\t& B0.5--B7\t\t&5156\t& B0.5--B7 \t\\\\ 4431\t& B1--B7\t\t&5272\t& B1.5--B5 \t\\\\ 5063\t& B1.5--B7\t\t&5282 \t& B2--B7 \t\\\\ 5086\t& B0.5--B7\t\t&5299 \t& B1.5--B4\t\\\\ 5127\t& B0.5--B7\t\t&5302\t& B1.5--B4\t\\\\ \\hline \\end{tabular} \\end{table} In Table \\ref{tab_touse} we list recommended Fe\\,{\\sc iii} lines which we believe, based on the present study, will provide reliable diagnostics for the iron abundance in early B-type stars. Lines have been selected based on the following criteria: \\begin{itemize} \\item Relatively strong, isolated feature, free from known blends. \\item Yields an abundance estimate within $\\pm$ 0.2 dex of the mean iron abundance for all well observed lines. \\end{itemize} \\begin{figure*} \\includegraphics[angle=0,width=0.8\\textwidth]{figure3_a.eps} \\caption{Observed Fe\\,{\\sc iii} lines in HD\\,79447 as listed in Table \\ref{tab_touse}, including the 5193 \\AA\\ line. Overplotted are theoretical fits (smooth line) to the Fe\\,{\\sc iii} lines, calculated using the average abundance estimate for the star of 7.63 dex.} \\label{fig_felines} \\end{figure*} The range of spectral types over which it is advisable to use the lines as an abundance diagnostic is also listed in Table \\ref{tab_touse}. This spectral type range is not the same for all lines, generally due to the fact that some of the weaker transitions are only detected over a restricted span of spectral types. Observations for all of the lines in this Table are shown in Fig. \\ref{fig_felines} for HD\\,79447. The line at 5193 \\AA\\ has been included in the Figure, but not the Table, because its derived abundance is more than 0.2 dex larger than the mean value for three of the stars studied here, namely HD\\,108002 (B1), HD\\,142768 (B1.5) and HD\\,53138 (B3). However, the feature, along with others observed (see Tables \\ref{tab_tcab}, \\ref{tab_sab} and \\ref{tab_mab}) may be suitable for diagnostic use, depending on the quality of the spectra used. We note that the stars included in this study have been selected due to being narrow lined and are typical of their individual luminosity classes. However, if objects with larger $\\upsilon$\\,sin{\\it i} values were employed, blending between Fe lines and other metal features may occur, thus, care should be taken when using objects with larger projected rotational velocities." }, "0710/0710.2145_arXiv.txt": { "abstract": "\\footnote{To appear in proceedings of the Puerto Vallarta Conference on ``New Quests in Stellar Astrophysics II: Ultraviolet Properties of Evolved Stellar Populations'' eds. M. Chavez, E. Bertone, D. Rosa-Gonzalez \\& L. H. Rodriguez-Merino, Springer, ASSP series.} Recent HST/ACS images of M82 covering the entire galaxy have been used to detect star clusters. The galaxy is known to contain a young population (age $<10^7$~yr) in its starburst nucleus, surrounded by a post-starburst disk of age $<10^9$~yr. We detect more than 650 star clusters in this galaxy, nearly 400 of them in the post-starburst disk. These data have been used to derive the luminosity, mass and size functions separately for the young nuclear, and intermediate-age disk clusters. In this contribution, we discuss the evolutionary status of these clusters, especially, on the chances of some of these clusters surviving to become old globular clusters. ", "introduction": "\\label{sec:1} Super star clusters (SSCs) and globular clusters (GCs) represent the youngest and the oldest stellar aggregates known in the Universe. The environments in which these two kinds of clusters are found are vastly different --- SSCs are found in violent star-forming regions, whereas GCs are found in the halos of galaxies. Yet, the similarity in their compactness and mass, is a reason compelling enough to think of an evolutionary connection between them. The growing popularity of the hierarchical model of galaxy formation in the years following the discovery of SSCs, and the possibility of observing the epochs of galaxy (and GC) formation at high redshifts, have also generated interest in looking for a common origin for these two seemingly different classes of clusters. In order to investigate the relation between the two types of clusters, it is important to analyze the survival of SSCs for a Hubble time. Star clusters are vulnerable to a variety of disruption processes that operate on three different timescales \\citep[see][for more details] {Fal01, Mai04, deG07}. On short timescales ($t\\sim10^7$~yr), the exploding supernovae and the resulting superwinds are responsible for cluster expansion and disruption, a process popularly dubbed as infant mortality. On intermediate timescales ($10^7 {\\rm few} \\times 10^8$~yr), stellar dynamical processes, especially evaporation due to two-body scattering, and tidal effects on a cluster as it orbits around the galaxy, known as gravitational shocks, come into play in the removal of stellar mass from clusters. The GCs represent those objects that have survived all these processes, whereas young SSCs are just experiencing them. Intermediate age SSCs are the ideal objects to investigate the influence of disruption processes on the survival of star clusters. Almost all the star formation in the disk of M82 took place in a violent disk-wide burst around 100--500~Myr ago, following the interaction of M82 with the members of M81 group \\citep{May06}. Cluster formation is known to be efficient during the burst phase of star formation \\citep{Bas05}, and hence we expect large number of clusters of intermediate age ($\\sim100$~Myr) in its disk. Hence, M82 offers an excellent opportunity to assess the evolutionary effects on the survival of star clusters, and to look for a possible evolutionary connection between the SSCs and GCs. ", "conclusions": "Luminosity and Mass functions of star clusters in M82 follow power-law functions, with the power law index showing a tendency for flattening of the profile with age. In other words, there is a deficiency of low-mass clusters among the older clusters. We also find the mean size of the older clusters to be smaller as compared to the younger clusters for masses $<10^5$~\\msun. These two results together imply the selective destruction of loose clusters. The tidal forces experienced by the clusters as they orbit around the galaxy lead to exactly such a destruction process. If this process continues in M82, the LF of surviving clusters can mimic the presently observed LF of the Galactic GCs, provided the clusters move around the galaxy in highly elliptical orbits, with perigalactic distance as small as 350~pc. The resulting LF contains 85 clusters with the function peaking at the same luminosity as for the Galactic GCs at 5~Gyr age, and fainter by $\\sim$0.5~mag at 10~Gyr. On the other hand, if the clusters move in nearly circular orbits, the LF will retain the power-law form, with the number of surviving clusters even higher. \\vspace*{0.3cm} This work is partly supported by CONACyT (Mexico) research grants 42609-F and 49942-F. We would like to thank the Hubble Heritage Team at the Space Telescope Science Institute for making the reduced fits files available to us." }, "0710/0710.0189_arXiv.txt": { "abstract": "I review the evolutionary connection between low-mass X-ray binaries (LMXBs) and pulsars with binary companions (bPSRs) from a stellar binary evolution perspective. I focus on the evolution of stellar binaries with end-states consisting of a pulsar with a low-mass ($<1.0 \\msun$) companion, starting at the point the companion's progenitor first initiates mass transfer onto the neutron star. Whether this mass transfer is stable and the physics driving ongoing mass transfer partitions the phase space of the companions's initial mass and initial orbital period into five regions. The qualitative nature of the mass-transfer process and the binary's final end-state differ between systems in each region; four of these regions each produce a particular class of LMXBs. I compare the theoretical expectations to the populations of galactic field LMXBs with companion-mass constraints and field bPSRs. I show that the population of accreting millisecond pulsars are all identified with only two of the four LMXB classes and that these systems do not have readily identifiable progeny in the bPSR population. I discuss which sub-populations of bPSRs can be explained by binary evolution theory and those that currently are not. Finally I discuss some outstanding questions in this field. ", "introduction": "Since the discovery of the class prototype \\citep{backer82}, there has been a posited evolutionary connection between millisecond radio pulsars (MSPs) and low-mass X-ray binaries (LMXBs)---mass transferring binaries with a neutron star (NS) accretor and donor companion with a mass $M_2 \\lesssim 1 \\msun$ \\citep{alpar82}. The central idea behind this connection is that LMXBs can provide the long-lived phase ($\\sim \\mathrm{Gyr}$) of moderate mass transfer rates ($\\Mdot \\lesssim \\Mdot_{\\mathrm{Edd}} \\approx 10^{-8} \\msun\\,\\mathrm{yr}^{-1}$, where $ \\Mdot_{\\mathrm{Edd}}$ is the Eddington mass-transfer rate) thought necessary to spin-up the NS to spin periods of $\\Pspin < 10 \\mathrm{ms}$ as observed in the MSP population. \\begin{figure} \\includegraphics[height=.425\\textheight]{proc_eps_sys_bPSRs_lmxbs.eps} \\caption{The galactic field population of radio pulsars with binary companions (filled circles) in the $M_2$-$\\Porb$ plane. The data are from the ATNF catalog \\citep{manchester05}. Plotted are each system's minimum $M_2$, with horizontal lines extending to the system's median $M_2$ (assuming $i$ is randomly distributed). Filled circle size indicates each pulsar's $\\Pspin$ (see plot legend). Symbols circumscribing a filled circle indicates the binary's eccentricity (again see plot legend). For comparison, LMXBs with independent $M_2$ estimates are plotted with open stars and filled triangles, the latter indicating the minimum $M_2$ for \\emph{accreting MSP} systems. For the open stars, the dash-dotted lines indicate the \\emph{total} estimated $M_2$ range in each system. For the accreting MSPs, the dash-dotted lines extend to the median $M_2$. This sample of LMXB systems represents a union of systems with $M_2$ estimates in the \\citet{ritter98} catalog and a targeted literature search on LMXB systems where $\\Pspin$ has been determined. Thus, this plot likely does not present in total the current census of LXMBs with $M_2$ estimates. } \\label{fig:bPSRs_sys} \\end{figure} Observational evidence for millisecond variability in LMXBs has been growing steadily in the form of detections of kilohertz quasi-periodic oscillations (kHz QPOs), X-ray burst oscillations, and accretion-powered oscillations \\citep[see][]{chakrabarty05}. In terms of support for the LMXB-MSP connection, pride of place has been given to the accretion-powered pulsations systems (also known as accreting MSPs) since the pulsation period in these systems are identifiable directly with $\\Pspin$ \\citep{chakrabarty05}. However, recent work has also shown that $\\Pspin$ is of order the oscillation periods in the kHz QPO and X-ray burst oscillation sources \\citep{chakrabarty03,strohmayer03,wijnands03,linares05}, establishing that these systems also harbor a rapidly rotating NS. Thus, it is now well established that NSs in LMXBs can be spinning rapidly enough to produce radio MSPs once the LMXB phase ends. However, understanding fully the evolutionary connection between LMXBs and MSPs requires not only explaining the NS's spin evolution but also accounting for other properties seen in the radio MSP population. In particular, this includes the distribution in orbital period, $\\Porb$ of MSPs that retain a remnant binary companion, how this remnant's $M_2$ correlates with $\\Porb$, the distribution of binary eccentricity, $e$, and the production of \\emph{isolated} MSPs. Indeed, the ideal test of accretion-torque theory would be accomplished by understanding the evolution in the LMXB-phase well enough to correlate final $\\Pspin$ with these other quantities. This is a rather ambitious goal since the $\\Pspin$ evolution depends not only on the secular $\\Mdot$ evolution but also on the efficiency with which matter accretes onto the NS, whether accretion onto the NS occurs sporadically due to disk instabilities, how $\\Pspin$ evolves during mass-transfer outbursts and when unstable disks are in quiescent phases, and how each LMXB transitions into an MSP system. Turning from where one would like to be to where we are now, my goal for this contribution is to approach the LMXB-MSP connection from the vantage point of stellar binary evolution theory. To do so, I will expand the view somewhat and review our understanding of NS-main sequence (MS) binaries whose evolutionary endpoints are NSs with a low-mass ($M_2 \\lesssim 1.0 \\msun$) binary companion. In doing so, the focus of the discussion will shift from solely the MSP population to making connections between NS-MS binaries and various populations of radio pulsars in binaries, bPSRs (one would like to simply say ``binary pulsars'' here, but the discovery of \\emph{the} Binary Pulsar \\citep{burgay03} has lead to this term often causing confusion; I'll leave it to the reader to decide whether the scientific windfall from this system compensates sufficiently for necessitating such unwieldy terminology). The starting point for this discussion is Figure \\ref{fig:bPSRs_sys}, which shows the location of galactic field bPSRs in the $\\Porb$-$M_2$ plane by the filled circles \\citep{manchester05}. Only field sources are included so as to compare theory to a sample of bPSRs whose properties have not been influenced by dynamic interactions. The filled circles indicate each bPSR's minimum-$M_2$; the horizontal lines extend to each system's median $M_2$ (corresponding to a binary inclination of $i = 60^\\circ$). Spin period and $e$ are encoded via filled circle size and circumscribed symbols, respectively. Filled triangles show the same information for accreting MSP systems \\citep{chakrabarty98,galloway02,galloway05,kaaret06,krimm07,markwardt02,markwardt03}. Open stars indicate other field LMXB systems with $M_2$ determinations \\citep{bhattachar06,casares06,cominsky89,cornelisse07,finger96,heinz01,hinkle06,jonker01,parmer86,pearson06,reynolds97}; for these latter systems, the dashed-dotted horizontal lines show the estimated range of $M_2$ (not its median value). In the following, I'll compare the predictions of binary evolution theory for how systems evolve in this $\\Porb$-$M_2$ plane to the location of systems in Fig. \\ref{fig:bPSRs_sys}. While this will allow tentative positive identifications of the evolutionary connections discussed above, it will also serve to highlight sub-populations of bPSRs whose formation is \\emph{not} currently explained by binary evolution theory. I'll discuss the evolution of NS-MS binaries, focusing on how initial conditions lead to four different classes of X-ray binaries. I'll compare between this theory and the observations, pointing out where agreement between the two is better and worse. Finally, I'll close by discussing several open questions related to the evolution of LMXBs and bPSR formation. For other discussions and reviews of the bPSR population see \\citep{lorimer05} and \\citep{phinney94}. Also see \\citep{flamb05} for a conference proceeding that reviews NS spin evolution under accretion. ", "conclusions": "" }, "0710/0710.5881_arXiv.txt": { "abstract": "We report observations of a radio burst that occurred on the flare star AD~Leonis over a frequency range of 1120-1620~MHz ($\\lambda\\approx$18--27~cm). These observations, made by the 305~m telescope of the Arecibo Observatory, are unique in providing the highest time resolution (1~ms) and broadest spectral coverage ($\\Delta \\nu/\\nu=0.36$) of a stellar radio burst yet obtained. The burst was observed on 2005 April 9. It produced a peak flux density of $\\sim 500$ mJy and it was essentially 100\\% right-circularly polarized. The dynamic spectrum shows a rich variety of structure: patchy emission, diffuse bands, and narrowband, fast-drift striae. Focusing our attention on the fast-drift striae, we consider the possible role of dispersion and find that it requires rather special conditions in the source to be a significant factor. We suggest that the emission may be due to the cyclotron maser instability, a mechanism known to occur in planetary magnetospheres. We briefly explore possible implications of this possibility. ", "introduction": "The use of radio dynamic spectra has played a central role in identifying and clarifying the physical mechanisms at work in the solar corona \\citep[see][for reviews]{1985srph}. The application of similar techniques to active stars has long been an important goal, but it has been hampered by limitations in available instrumentation. Past studies of radio emission from M dwarf flare stars led to the discovery of extreme stellar radio bursts, characterized by close to 100\\% circularly polarized emission with brightness temperatures in excess of 10$^{14}$K and durations less than a few tens of milliseconds \\citep{gudel1989,bastian1990}. However, these spectroscopic investigations of the coherent radio bursts on flare stars have typically been limited by relatively long integration times \\citep{bb1987,gudel1989} and/or limited frequency bandwidth ratios $\\Delta\\nu/\\nu$ \\citep[usually just a few percent; e.g.,][]{bastian1990, abada1997a}. The necessary combination of high time resolution and a large frequency bandwidth ratio has only been available infrequently \\citep{stepanov2001, zaitsev2004}, precluding measurements of key parameters such as the intrinsic frequency bandwidth or frequency drift rate of the radio bursts, making the interpretation of these puzzling events difficult. It is only with the recent advent of radio spectrometers capable of supporting both a large bandwidth ratio and high time resolution simultaneously that progress in understanding the physics of radio bursts in the coronas of other stars becomes possible. AD~Leonis, a young disk star at a distance of 4.9 pc from the Sun, is one of the most active flare stars known, producing intense, quasi-steady chromospheric and coronal emissions \\citep{hawleyetal2003, hunschetal1999, jackson1989} seen at UV, X-ray, and radio wavelengths. The star is also highly variable, producing flares from radio to X-ray wavelengths \\citep[e.g.,][]{bastian1990, hawleypettersen1991, hawleyetal2003, hawleyetal1995, favataetal2000}. Its propensity for frequent and extreme radio bursts \\citep[with intensities peaking at $>$ 500 times the quiescent radio luminosity of 5.5$\\times$10$^{13}$ erg s$^{-1}$ Hz$^{-1}$;][]{jackson1989} makes it a frequent target for radio investigations of stellar flares. In a previous paper \\citep[][hereafter, Paper I]{ob2006} we described the initiation of a pilot program to observe active M dwarfs with the Arecibo Observatory's Wideband Arecibo Pulsar Processor (WAPP), and first results from that program. Here we describe the next phase, which increased the time resolution by a factor of 10 to 1~ms. ", "conclusions": "We have described observations of a unique set of stellar radio bursts, which take advantage of the wide bandwidth and high time resolution capabilities of the Wideband Arecibo Pulsar Processor at the Arecibo Observatory. These ultra-high time resolution observations reveal phenomena that differ from those previously described using a similar observational setup, pointing out the complexity and diversity of processes likely occurring in stellar coronal plasmas. Whereas in Paper I we concluded that a plasma emission process appeared to be producing the two types of radio bursts observed in June 2003, in the current paper we prefer a different explanation, a cyclotron maser instability, for the fast-drift striae observed in April 2005. While all sets of phenomena show drifting structures of highly circularly polarized radiation, key discriminants between them are the durations and bandwidths of spectral features, as well as the magnitude and sign of the drift rates. In Paper I and here, we have demonstrated that the analysis of dynamic spectra of stellar radio bursts provide observational constraints which can be used as a measure to gauge the likelihood that a particular emission process is operative. Extensions of the current observational setup can look for dynamics at even higher time resolution, search for harmonic emissions over larger frequency bandwidths, expand the observational program to other dMe flare stars, and search for high time resolution behavior on other classes of active stars. Given the complexity of solar radio emissions at meter wavelengths compared with the already rich variety of decimetric phenomena, the observational results presented here for the dMe flare star AD Leo suggest that the next generation of radio instrumentation, particularly at metric wavelengths, promises to reveal a wealth of new phenomena which can diagnose plasma processes occurring in stellar coronae. As highly circularly polarized radio emission appears to be a common phenomenon on active stars, these spectacular radio bursts on M dwarf flare stars apparently represent the tip of the iceberg of stellar coronal plasma physics soon to be available for study." }, "0710/0710.5286_arXiv.txt": { "abstract": "We present new observations of the strongly-barred galaxy NGC~1365, including new photometric images and Fabry-Perot spectroscopy, as well as a detailed re-analysis of the neutral hydrogen observations from the VLA archive. We find the galaxy to be at once remarkably bi-symmetric in its I-band light distribution and strongly asymmetric in the distribution of dust and in the kinematics of the gas in the bar region. The velocity field mapped in the \\ha\\ line reveals bright HII regions with velocities that differ by 60 to $80\\;$\\kms\\ from that of the surrounding gas, which may be due to remnants of infalling material. We have attempted hydrodynamic simulations of the bar flow to estimate the separate disk and halo masses, using two different dark matter halo models and covering a wide range of mass-to-light ratios ($\\Upsilon$) and bar pattern speeds ($\\Omega_p$). None of our models provides a compelling fit to the data, but they seem most nearly consistent with a fast bar, corotation at $\\sim1.2r_B$, and $\\Upsilon_ I \\simeq 2.0 \\pm 1.0$, implying a massive, but not fully maximal, disk. The fitted dark halos are unusually concentrated, a requirement driven by the declining outer rotation curve. ", "introduction": "The centrifugal balance of the circular flow pattern in a near-axisymmetric spiral galaxy yields a direct estimate of the central gravitational attraction as a function of radius. However, the division of the mass giving rise to that central attraction into separate dark and luminous parts continues to prove challenging. The radial variation of the circular speed simply does not contain enough information to allow a unique decomposition between the baryonic mass, which has an uncertain mass-to-light ratio, $\\Upsilon$, and the dark halo, whose density profile is generally described by some adopted parametric function \\citep{albada3198, LF89, BSK04}. Predictions for $\\Upsilon$ from stellar population synthesis models that match broad-band colors \\citep[e.g.][]{Bell03} are useful, but not precise. Despite intense effort, they are still sufficiently uncertain to be consistent with both maximum and half-maximum disk, which is the range of disagreement \\citep[e.g.][]{Sackett97, Bottema97, S99Rutgers}. \\citet{McGaugh05} argues that the values can be refined by minimizing the scatter in the Tully-Fisher and/or mass discrepancy-acceleration relation. A number of dynamical methods have been employed to break the disk-halo degeneracy. \\citet{Casertano83}, \\citet{bosma98}, and others have suggested that the slight decrease in orbital speed near the edge of the optical disk of a bright galaxy -- the ``truncation signature'' -- could be used as an indicator of disk $\\Upsilon$, but in practice it does not provide a tight constraint. \\citet{ABP87} and \\citet{Fuchs03} attempt to constrain the disk mass using spiral structure theory. \\citet{Bottema97} and \\citet{verheijen04} measure the vertical velocity dispersion of disk stars in a near face-on galaxy, which they assume has the same mean thickness of similar galaxies seen edge-on \\citep{KvdKdG}, to constrain the disk mass. A similar approach is reported by \\citet{CD04} using velocity measurements of individual planetary nebulae. One of the most powerful, although laborious, methods for barred galaxies was pioneered by Weiner, Sellwood \\& Williams (2001), who made use of the additional information in the driven non-circular motions caused by the bar. By modeling the observed non-axisymmetric flow pattern of the gas in a 2-D velocity map, they were able to determine the mass-to-light ratio of the visible disk material. They found that the luminous disk and bar contributed almost all the central attraction in NGC~4123 inside $\\sim 10\\;$kpc, requiring the dark halo to have a very low central density. \\citet{Ben3095} reports a similar result for a second case, NGC 3095. The method has also been applied by \\citet{perez04} for several barred galaxies and by \\citet{kranz03} who modeled motions caused by spiral arms. \\citet{BEG03} present a similar study for the Milky Way. Earlier studies \\citep[\\eg][]{DA83} did not attempt to separate the disk from the dark matter halo \\citep[see][for a review]{SW93}. Here, we apply the \\citep{wein2} method to the more luminous barred galaxy NGC~1365 in the Fornax cluster. As one of the most apparently regular, nearby barred spiral galaxies in the Southern sky, NGC~1365 was selected by the Stockholm group for an in-depth study \\citep[see \\eg][]{Lindblad1365}. Hydrodynamic models of the bar flow pattern were already presented by \\citep{Linlinatha}, based mainly on the velocities of emission-line measurements from many separate long-slit observations. J\\\"ors\\\"ater \\& van Moorsel (1995, hereafter JvM95) present a kinematic study using the 21 cm line, which suggests that the galaxy is somewhat asymmetric in the outer parts, where the shape of the rotation curve is hard to determine. \\citet{sandqvist} find substantial amounts of molecular gas, but only within 2~kpc of the nucleus, which is resolved in interferometric observations \\citep{Sakamoto07} into a molecular ring in the center plus a number of CO hot spots. \\citet{Galliano05} have found previously unknown MIR sources in the inner 10\\arcsec\\ around the AGN. They are able to correlate some of these MIR sources with radio sources, which they interpret in terms of embedded star clusters because of the lack of strong optical counterparts. \\citet{JCA97} present H-band photometry of the bright inner disk, finding an elongated component in the central region suggesting that the NGC~1365 is a double-barred galaxy, although they note that the light in this component is not as smooth as in their other nuclear bar cases. \\citet{LSKP} also classify it as a double barred galaxy. However, \\citet{Emsellem01} and \\citet{Erwin04} argue against a nuclear bar, citing an HST NICMOS image which resolves the feature into a nuclear spiral. \\citet{Emsellem01} also present stellar kinematics from slit spectra using the $^{12}$CO bandhead. They propose a model for the inner 2.5 kpc of NGC 1365 consisting of a decoupled nuclear disk surrounded by spiral arms within the inner Lindblad resonance (ILR) of the primary bar. \\citet{Beck05} observed NGC 1365 in radio continuum at 9\\arcsec-25\\arcsec\\ resolution and find radio ridges roughly overlapping with the dust lanes in the bar region. They propose that magnetic forces can control the flow of gas at kiloparsec scales. Here we present new photometric images, a full 2-D velocity map of the \\ha\\ emission, and a reanalysis of the neutral hydrogen data from JvM95. We also compare many hydrodynamic models to these new data in an effort to determine the separate disk and dark halo masses in this galaxy. ", "conclusions": "We have presented a detailed study of the strongly-barred galaxy NGC~1365, including new photometric images and Fabry-Perot spectroscopy, as well as a detailed re-analysis of the neutral hydrogen observations by J\\\"ors\\\"ater \\& van Moorsel (1995). We find the galaxy to be at once both remarkably bi-symmetric in its I-band light distribution and strongly asymmetric in the distribution of dust and gas, and in the kinematics of the gas. These asymmetries extend throughout the galaxy, affecting the bar region, the distribution of gas in the spiral arms and the neutral hydrogen beyond the edge of the bright disk. The velocity field mapped in the \\ha\\ line showed bright HII regions with velocities that differed by up to $\\sim 80\\;$\\kms\\ from that of the surrounding gas. Our sparsely-sampled line profiles in these anomalous velocity regions hint at unresolved substructure, suggesting a possible double line profile. The strong bar and spiral arms complicate the determination of the projection geometry of the disk, assuming it can be characterized as flat in the inner parts. The inclination of the plane we derive from the kinematic data is smaller by about 10$^\\circ$ from that determined from the photometry. The strong spiral arms that cross the projected major axis far out in the disk seem likely to bias the photometric inclination and we therefore adopt, in common with other workers, the inclination derived from the gas kinematics. This preference is supported by the much poorer fits to the observed kinematics obtained when we adopt the photometric inclination (\\S~\\ref{comparsect}). Our attempts to derive the rotation curve of NGC~1365 were complicated by the fact that neither the \\ha\\ nor the HI velocity maps are consistent with a simple circular flow pattern over a significant radial range. The bar and spirals clearly distort the gas flow in the luminous disk. The neutral hydrogen extends somewhat beyond the visible disk but unfortunately has neither a uniform distribution nor regular kinematics. JvM95 attempted to fit a warp to the outer HI layer that extends into the visible disk, and derived a strongly declining rotation curve. We chose instead to assume a coplanar flow out to a deprojected radius of 255\\arcsec\\ and to neglect the asymmetric velocities in the neutral hydrogen beyond. The velocities derived separately from the \\ha\\ and HI data are in good agreement. Our resulting rotation curve shows a gentle decrease beyond a radius of $\\sim 10\\;$kpc, similar to those observed in other massive galaxies \\citep{CvG91, noordermeer07}. We used our deprojected I-band image to estimate the gravitational field of the luminous matter, which can be scaled by a single mass-to-light ratio, $\\Upsilon_I$. We also employed a gradual increase to $\\Upsilon_I$ in the central few kpc to allow for an older bulge-like stellar population, although the light distribution does not appear to have a substantially greater thickness near the center. We combined the central attraction of the axially-symmetrized disk for various values of $\\Upsilon_I$ with two different halo models to fit the observed rotation curve in the region outside the bar -- finding, as always, no significant preference for any $\\Upsilon_I$. We attempt to fit hydrodynamic simulations of the gas flow pattern in the bar region, in order to constrain $\\Upsilon_I$. For each type of halo adopted, we run a grid of simulations covering a range of both $\\Upsilon_I$ and $\\Omega_p$, the pattern speed of the bar. We then project each simulation to our adopted orientation of the galaxy and compare the gas flow velocities in the model with those observed. Since the light distribution in NGC~1365 is highly symmetric, our simulations were constrained to be bi-symmetric, yet the observed gas flow has strong asymmetries. None of our simulations is capable, therefore, of fitting both sides of the bar simultaneously. The anomalous position of the dust lane in the western part of the bar suggests that side is the more likely to be disturbed, and we therefore fit our models to the eastern half of the bar only. After smoothing the model to match the resolution of the kinematic data and masking out five blobs of gas with strongly anomalous velocities, we are able to obtain moderately satisfactory fits to the remaining velocities. The best fit pattern speed is $\\Omega_p = 24\\;$\\kms$\\rm{kpc}^{-1}$ for both types of halo which places corotation for the bar at $r_L \\simeq 1.23r_B$, in excellent agreement with the value found by \\citet{Linlinatha} and consistent with most determinations of this ratio for other galaxies \\citep[\\eg][]{ADC03}. Our estimated mass-to-light ratio values are $\\Upsilon_I \\simeq 2.50 \\pm 1$ for the isothermal halo models and $\\Upsilon_I=1.75 \\pm 1$ for the NFW halo. While the constraints are disappointingly loose, the preferred mass-to-light ratio in both our halo models, $\\Upsilon_I \\simeq 2.0 \\pm 1$, is consistent with that obtained by \\citet{wein2} and \\citet{Ben3095} in two other cases. For NGC~1365, however, this value implies a massive, but not fully maximal, disk, and we do not find support for the disk-only model with no halo that was suggested by JvM95. Although such a model can reproduce the declining rotation curve (see their Fig. 24), the simulated gas flow produced by such a model (which in the I band is $\\Upsilon\\sim 3.75$) is quite strongly excluded. The preferred value of $\\Upsilon_I$ is nicely consistent with those obtained in the two previous studies using this method \\citep{wein2, Ben3095}, but suggest somewhat more massive disks than predicted by population synthesis models \\citep{BdJ,Bell03} for galaxies of these colors. The halos of our two models required to fit the declining rotation curve in the outer disk are distinctly non-standard, however. The circular speed in the pseudo-isothermal model declines steadily outside the large core, while the NFW halo has a very high concentration and small scale radius. Even allowing for compression of the halo as the massive disk forms within it, the original NFW halo has $c \\simeq 22$, $V_{200} \\simeq 123\\;$kms. The total dark matter mass out to $r_{200}$ in this case is less than three times our estimated disk mass, and the halo is quite unlike those predicted by LCDM models for a galaxy of this mass. The disturbed distribution and kinematics of the gas in this galaxy clearly complicates our attempt to identify a preferred mass model. Its projected position near the center of the Fornax cluster, together with its velocity within 200~\\kms\\ of the cluster mean, suggest it is a cluster member. The disturbed nature of the outer HI distribution should not therefore be regarded as surprising. But the high central density of this massive galaxy should ensure that tidal forces have little influence on the inner part, where we indeed see that the bar and inner spirals are very pleasingly bi-symmetric in the I-band light. The existence of such strong asymmetries in the inner parts of the gas and dust is rather surprising, therefore. The asymmetry in the dust distribution and the kinematic map, combined with the existence of a number of patches of \\ha\\ emission with anomalous velocities all suggest that the agent that caused the disturbance was an infalling gas cloud. We cannot say whether the gas was an isolated intergalactic cloud not associated with a galaxy, or whether it could be a stream of debris from a gas-rich dwarf galaxy that had been tidally disrupted. The anomalous velocities clearly suggest that the infalling gas has yet to be assimilated in the disk of NGC~1365." }, "0710/0710.5765_arXiv.txt": { "abstract": "We examine halo gas cross sections and covering fractions, $f_c$, of intermediate redshift {\\MgII} absorption selected galaxies. We computed statistical absorber halo radii, $R_{\\rm x}$, using current values of $dN/dz$ and Schechter luminosity function parameters, and have compared these values to the distribution of impact parameters and luminosities from a sample of 37 galaxies. For equivalent widths $W_r(2796) \\geq 0.3$~{\\AA}, we find $43 \\leq R_{\\rm x} \\leq 88$~kpc, depending on the lower luminosity cutoff and the slope, $\\beta$, of the Holmberg--like luminosity scaling, $R \\propto L^{\\beta}$. The observed distribution of impact parameters, $D$, are such that several absorbing galaxies lie at $D>R_{\\rm x}$ and several non--absorbing galaxies lie at $D < R_{\\rm x}$. We deduced $f_c$ must be less than unity and obtain a mean of $\\left< f_c \\right> \\sim 0.5$ for our sample. Moreover, the data suggest halo radii of {\\MgII} absorbing galaxies do not follow a luminosity scaling with $\\beta$ in the range of $0.2-0.28$, if $f_c= 1$ as previously reported. However, provided $f_c \\sim 0.5$, we find that halo radii can remain consistent with a Holmberg--like luminosity relation with $\\beta \\simeq 0.2$ and $R_{\\ast} = R_{\\rm x}/\\sqrt{f_c} \\sim 110$~kpc. No luminosity scaling ($\\beta=0$) is also consistent with the observed distribution of impact parameters if $f_c \\leq 0.37$. The data support a scenario in which gaseous halos are patchy and likely have non--symmetric geometric distributions about the galaxies. We suggest halo gas distributions may not be govern primarily by galaxy mass/luminosity but also by stochastic processes local to the galaxy. ", "introduction": "Understanding galaxy formation and evolution is one of the most important topics of modern astronomy. The extended distribution of baryonic gas surrounding galaxies holds great potential for constraining theories of their formation. However, the sizes of gaseous galaxy halos along with the distribution of gas within are not well understood. Numerical models have been able to synthesize the formation and evolution of large scale structures, however, there are unresolved issues regarding the evolution of individual galaxies and halos. The halo baryon--fraction problem \\citep[e.g.,][]{mo02} and the rapid cooling of gas \\citep[e.g.,][]{white78} result in galaxy halos which have little or no gas soon after they form. These effects are not seen in the observable universe since there is an abundance of galaxies where gas has been detected in halos via quasar absorption lines. From an observational standpoint, quasar absorption lines provides a unique means of probing the extent and abundance of halo gas. Although, quasar absorption line observations to date are sufficient the recognize the aforementioned problems, they are lacking the detail required to statistically constrain the distribution of the baryonic gas in the halos of simulated galaxies. Cross--correlations between absorbers and galaxies hold the promise to yield useful information on cloud sizes and halo gas covering fractions. First steps towards incorporating multi--phase gas in semi--analytical models and numerical simulations suggest that warm gas in halos extends out to galactocentric distances of $\\sim 150$~kpc with cloud covering fractions of $\\sim 0.25-0.6$ \\citep{maller04,kaufmann06}. The association of {\\MgIIdblt} doublet absorption in quasar spectra with normal, bright, field galaxies has been firmly established \\citep[e.g.,][]{bb91,sdp94,cwc-china}. In an effort to understand halo sizes and gas distributions, \\citet[][hereafter S95]{steidel95} searched for foreground galaxies associated with {\\MgII} absorption within $\\sim10''$ ($\\sim65$~kpc for $z=0.5$) of quasars\\footnote{Throughout we adopt a $h=0.70$, $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda}=0.7$ cosmology. All quoted physical quantities from previously published works have been converted to this cosmology.}. The sample consisted of 53 absorbing and 14 non--absorbing galaxies with a {\\MgII} $\\lambda 2796$ equivalent width sensitivity limit of $W_{r}(2796) > 0.3$~{\\AA}. S95 directly fitted the data by assuming a Holmberg--like luminosity scaling, \\begin{equation} R(L) = R_{\\ast} \\left( \\frac{L}{L^{\\ast}} \\right) ^{\\beta} \\quad {\\rm kpc}, \\label{eq:rl} \\end{equation} and minimizing the number of non--absorbing and absorbing galaxies above and below the $R(L)$ relation. The best fit obtained clearly showed that absorbing and non--absorbing galaxies could be separated and that the halo radii $R(L_K)$ and $R(L_B)$ scale with luminosity with $\\beta = 0.15$ and $\\beta = 0.2$, respectively, where an $L_B^{\\ast}$ galaxy has a gas halo cross section of $R_{\\ast} = 55$~kpc. Furthermore, since almost none of the absorbing galaxies were observed above the $R(L)$ boundary and that almost none of the non--absorbing galaxies were observed below the $R(L)$ boundary, S95 inferred that {\\it all\\/} $L> 0.05L^{\\ast}$ galaxies are hosts to {\\MgII} absorbing gas halos characterized by a covering fraction of unity and a spherical geometry which truncates at $R(L)$. Examination of this now ``standard model'' has been the subject of several theoretical studies \\citep[e.g.,][]{cc96,mo96,lin01}. \\citet{gb97} determined a steeper value of $\\beta = 0.28$ for the B--band luminosity obtained from a best fit to the upper envelope of the distribution of impact parameters of 26 absorbing galaxies. They found $R_{\\ast} = 67$~kpc. Using a reverse approach of establishing foreground galaxy redshifts and then searching for {\\MgII} absorption in the spectra of background quasars yields results inconsistent with a covering fraction of unity. For example, \\citet{bowen95} identified 17 low--redshift galaxies with background quasar probing an impact parameter range between $3-162$~kpc. Galaxies that were probed at impact parameters greater than 13~kpc had no absorption in the halo ($W_{r}(2796) \\geq 0.40-0.9$~{\\AA}), however, four of the six galaxies within 13~kpc of the halo produced {\\MgII} absorption. For intermediate redshift galaxies, \\citet{bechtold92} reported a covering fraction $f_c \\simeq 0.25$ for $W_{r}(2796) \\geq 0.26$~{\\AA} for eight galaxies with $D \\leq 85$ kpc. Also, \\citet{tripp-china} reported $f_c \\sim 0.5$ for $W_{r}(2796) \\geq 0.15$~{\\AA} for $\\sim 20$ galaxies with $D \\leq 50$ kpc. These results are also consistent with the findings of \\citet{cwc-china} who reported very weak {\\MgII} absorption, $W_{r}(2796) < 0.3$~{\\AA}, well inside the $R(L)$ boundary of bright galaxies; these galaxies would be classified as ``non--absorbers'' in previous surveys. They also report $W_r(2796) > 1$~{\\AA} absorption out to $\\simeq 2 R(L)$. All these results suggest that there are departures from the standard model, that the covering fraction of {\\MgII} absorbing gas is less than unity, and that the halo sizes and the distribution of the gas appear to diverge from the $R(L)$ relation with spherical geometry. Another approach to understanding halo sizes and gas distributions is to determine the statistical properties of {\\MgII} absorbing gas and then compute the statistical cross section from the redshift path density, $dN/dz$ \\citep[see][]{lanzetta95}. The downfall of this method is that a galaxy luminosity function must be adopted in order to estimate $R_{\\ast}$. \\citet{nestor05} acquired a sample of over 1300 {\\MgII} absorption systems, with $W_{r}(2796) \\geq 0.3$~{\\AA} from the Sloan Digital Sky Survey (SDSS). Using the $K$--band Holmberg--like luminosity scaling and luminosity function of MUNICS \\citep{drory03}, Nestor {\\etal} computed $R_{\\ast} = 60-100$~kpc for adopted minimum luminosity cutoffs of $L_{min}=0.001-0.25L^{\\ast}$. They found no redshift evolution of $R_{\\ast}$ over the explored range of $0.3\\leq z \\leq 1.2$. \\citet{zibetti06} studied the statistical photometric properties of $\\sim2800$ {\\MgII} absorbers in quasar fields imaged with SDSS. Using the method of image stacking, they detected low--level surface brightness (SB) azimuthally about the quasar. The SB profiles follow a decreasing power law with projected distance away from the quasar out to $100-200$~kpc. These results imply that absorption selected galaxies may reside out to projected distances of 200~kpc. However, it is worth noting that the extended light profiles may be an artifact of clustering of galaxies. Cluster companions of the {\\MgII} absorbing galaxies could extend the observed light profile over hundreds of stacked images. Thus, one would infer that {\\MgII} absorbing galaxies are present at a larger impact parameters than would be found in direct observation of individual galaxies. Motivated by recent expectations from simulations that halo gas is dynamically complex and sensitive to the physics of galaxy formation, we investigate the standard halo model of {\\MgII} absorbers. We also aim to provide updated constraints on $f_c$ and $\\beta$ for galaxy formation simulations. In this paper, we demonstrate that $f_c < 1$ and question the validity of the Holmberg--like luminosity scaling (Eq.~\\ref{eq:rl}). Using high resolution quasar spectra, we explore {\\MgII} absorption strengths to an order of magnitude more sensitive than previous surveys which allow us to re--identify non--absorbing galaxies as ``weak'' absorbing galaxies. In \\S~\\ref{sec:data} we describe our sample and analysis. In \\S~\\ref{sec:results}, we present new calculations of the statistical absorber radius computed using the statistically measured absorption path density $dN/dz$ and the Schechter luminosity function. We then compare these values to the empirical results of S95 and to a sample of known {\\MgII} absorption selected galaxies with measured luminosities and impact parameters. We also examine how individual halos behave with respect to the statistical halo. In \\S~\\ref{sec:dis}, we discuss the properties and distribution of gas in halos. Our concluding remarks are in \\S~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} In conclusion, the gas covering fraction must be less than unity since the observed impact parameter distribution of absorbing galaxies does not fall exclusively within the statistical absorber halo radius in the range of $43 \\leq R_{\\rm x}\\leq 88$~kpc. The fact that some absorbing galaxies are found at $D>R_{\\rm x}$ and some non--absorbing galaxies are found at $D \\sim 0.5$. It is possible that $f_c$ exhibits both a radial and an equivalent width dependence, though we cannot address this with our sample. Values of $f_c$ are likely to depend on galaxy star formation rates, and galaxy--galaxy mergers and harassment histories; processes that give rise to patchy and geometrically asymmetric gas distributions. Alternatively, the absorption properties of intermediate redshift halos may be governed by the dark matter over density, $\\Delta \\rho /\\rho$, and redshifts at which the galaxies formed \\citep{cwc06}. Our results also show that, if $f_c<1$, the sizes of {\\MgII} absorbing halos can still follow a Holmberg--like luminosity relation with $\\beta$ in the range of $0.2-0.28$ \\citep[S95;][]{gb97}, which corresponds to $R_{\\ast}\\sim 110$~kpc. If $\\beta=0$ is assumed, then $f_c \\leq 0.37$ for our sample to be consistent with no luminosity scaling. In semi--analytical models in which {\\MgII} absorbing gas is infalling and is pressure confined within the cooling radius of hot halos \\citep[e.g.,][]{mo96,burkert00,lin00,maller04}, a Holmberg--like luminosity relation in quasar absorption line systems naturally arises \\citep{mo96}. However, these models have great difficulty explaining {\\MgII} absorption at impact parameters greater than $\\sim 70$~kpc. If on the other hand halo gas spatial distributions are governed by stochastic mechanical processes, as suggested by \\citet{kacprzak07}, then there is no {\\it a priori} reason to expect a clean halo--size luminosity scaling. It is likely that some combination of these scenarios contribute to the statistical values of $f_c$ and $\\beta$. Thus, it is reasonable to suggest that {\\MgII} halos sizes may not be strictly coupled to the host galaxy luminosity. Further work on the cross--correlations between absorbers and galaxies would provide better estimates of $f_c$ and $\\beta$, two quantities that provide direct constraints of galaxy formation simulations. Also needed are additional constrains on the relative kinematics of the absorbing halo gas and galaxies \\citep[e.g.,][]{s02,ellison03,kacprzak07b}. What is required is the development of techniques to quantitatively compare observational data with mock quasar absorption line analysis of simulated galaxy halos \\citep{cwcaas06}." }, "0710/0710.5179.txt": { "abstract": "We propose a scenario for the formation of the Main Belt in which asteroids incorporated icy particles formed in the outer Solar Nebula. We calculate the composition of icy planetesimals formed beyond a heliocentric distance of 5 AU in the nebula by assuming that the abundances of all elements, in particular that of oxygen, are solar. As a result, we show that ices formed in the outer Solar Nebula are composed of a mix of clathrate hydrates, hydrates formed above 50 K and pure condensates produced at lower temperatures. We then consider the inward migration of solids initially produced in the outer Solar Nebula and show that a significant fraction may have drifted to the current position of the Main Belt without encountering temperature and pressure conditions high enough to vaporize the ices they contain. We propose that, through the detection and identification of initially buried ices revealed by recent impacts on the surfaces of asteroids, it could be possible to infer the thermodynamic conditions that were present within the Solar Nebula during the accretion of these bodies, and during the inward migration of icy planetesimals. We also investigate the potential influence that the incorporation of ices in asteroids may have on their porosities and densities. In particular, we show how the presence of ices reduces the value of the bulk density of a given body, and consequently modifies its macro-porosity from that which would be expected from a given taxonomic type. ", "introduction": "In recent years, some objects within the Main Belt of asteroids have been found to display cometary characteristics (Hsieh \\& Jewitt 2006). Objects such as 133P/Elst-Pizarro, P/2005 U1 and 118401 (1999 RE$_{70}$) occupy orbits that are entirely decoupled from Jupiter within the Main Belt, and are probably bodies that have undergone a recent collision, revealing previously buried volatile material, and leading to the observed dusty outgassing. In addition, present-day surface water ice and possible water sublimation have been reported on Ceres (Lebofsky et al. 1981; A'Hearn \\& Feldman 1992; Vernazza et al. 2005). This is consistent with recent Hubble Space Telescope (HST) observations which suggest that Ceres' shape is the result of the dwarf planet consisting of a rocky core surrounded by an ice-rich mantle (Thomas et al. 2005) - an idea in agreement with several thermal evolution scenarios (McCord \\& Sotin 2005) that suggest that the ice content of the asteroid is between 17\\% and 27\\%, by mass. These observations are supported by the evidence of hydrated minerals in meteorites which provide samples of rock from asteroids in the Main Belt. Most of these minerals formed as a result of water ice accreting with the chondritic meteorite parent bodies, melting, and driving aqueous alteration reactions (Clayton \\& Mayeda 1996; Jewitt et al. 2007). It seems likely, then, that some objects in the asteroid belt have incorporated significant amounts of water ice (and possibly other volatiles) during their formation in the early stages of the solar System. These bodies would have incorporated icy particles\\footnote{By icy particles is meant planetesimals composed of a mix of ices and rocks.} coming from the outer nebula that survived their inward drift due to gas-drag through the disk (Mousis \\& Alibert 2005 -- hereafter MA05). The volatile fraction incorporated in this manner could vary depending on the inward flux of icy planetesimals from the external region and the heliocentric location of the asteroid, together with the density of the proto-solar nebula. Given that the current asteroid belt lies closer to the Sun than the ``snow-line'', postulated to lie at around 5 AU in the Solar Nebula, these results are a little unexpected. In this context, understanding how volatiles were incorporated into the asteroids is therefore important, not only for the study of the asteroids themselves, but also for our understanding of the processes by which the solar system came into being. To this end, MA05 studied the possibility of determining the nature and composition of the ices which were incorporated into Ceres. They used a time dependant model of the Solar Nebula and showed that icy particles of sizes between 0.1 and 10 metres could drift from heliocentric distances greater than 5 AU to the present location of Ceres without encountering temperatures or pressures high enough to vaporise the ices within. The authors then suggested that ices produced in the outer Solar Nebula were transported inwards to become incorporated in the solids which accreted to form Ceres. The present work aims to improve upon the calculation detailed in MA05, along with expanding the results to involve the entire asteroid belt, rather than just its largest member. In particular, MA05 postulated that all volatiles, except CO$_2$\\footnote{CO$_2$ is the only major volatile species which does not form a clathrate hydrate in the Solar Nebula because it condenses as a pure ice prior to being trapped by water.}, were trapped by water in the form of hydrates or clathrate hydrates in the outer solar nebula. This assumption was supported by the work of Hersant et al. (2001) who estimated that Jupiter was formed at temperatures higher than $\\sim$40 -- 50 K. The accretion of ices in the form of hydrates and clathrate hydrates was thus required during the formation of the planet in order to explain the volatile enrichments observed in its atmosphere\\footnote{The abundances of volatile species in Jupiter's atmosphere have been measured using the mass spectrometer on board the {\\it Galileo} probe. These measurements reveal that the giant planet's atmosphere is enriched by a factor of $\\sim$3 in Ar, Kr, Xe, C, N, and S compared to the solar abundances (Owen et al. 1999).} (Gautier et al. 2001a,b). Indeed, since these ices are usually formed at temperatures higher than that reached by the nebula at the time of Jupiter's completion, as defined by Hersant et al. (2001), they can be incorporated in the planetesimals accreted by the giant planet during its growth. However, the amount of water that would be required in the nebula to trap all these volatiles as hydrates and clathrate hydrates exceeds that derived from the solar oxygen abundance. Therefore, MA05 made the {\\it ad hoc} hypothesis that oxygen was ``oversolar'' in the gas-phase in order to provide enough available water in the Solar Nebula\\footnote{ The oxygen abundance required for to allow the trapping of all volatile species in the form of hydrates or clathrate hydrates is $\\sim$1.9 times the solar abundance, with CO$_2$:CO:CH$_4$ = 1:1:1 and N$_2$:NH$_3$ = 1:1 (the nominal nebula gas phase ratios used in this work).}. Additionally, Hersant et al. (2001) only used an evolutionary Solar Nebula model to derive the disk's temperature at the time when the mass of Jupiter's feeding zone was equal to that of the gas in its current envelope. They thus neglected many important effects such as the influence of protoplanet formation on the structure of the disk (e.g. Fig. 2 of Alibert et al. 2004). However, recent giant planet core-accretion formation models that include migration, disk evolution, such as that proposed by Alibert et al. (2004), have shown that the disk's temperature can be as low as $\\sim$10 -- 20 K at the end of Jovian formation. This implies that Jupiter itself can accrete ices during its formation that were produced at temperatures lower than those required for clathration. As a result, no extra water is required in the nebula to allow all the volatile species to be trapped in clathrate hydrates, and the oversolar oxygen abundance condition in the nebula can be relaxed. In Section 2, we calculate the composition of ices produced in the outer Solar Nebula under the assumption that the abundances of all elements, in particular that of oxygen, are solar. In Section 3, we consider the inward migration of particles produced at various locations in the nebula, and at different times. This allows us to examine whether some planetesimals formed in the outer Solar Nebula may have drifted to the current position of the Main Belt without encountering temperature and pressure conditions high enough to vaporize the ices they contain. In Section 4, we examine the uncertainties in the determination of the physical properties of asteroids. We also investigate the potential influence that the incorporation of ices in these objects may have on their porosities and densities. Section 5 is devoted to summary and discussion. ", "conclusions": "In order to explain the presence of hydratation and cometary features in the Main Belt, we have proposed that asteroids incorporated during their formation icy particles formed in the outer Solar Nebula. We have then calculated the composition of the ices trapped in these planetesimals formed beyond a heliocentric distance of 5 AU in the nebula, in a manner consistent with the formation of Jupiter, by assuming that the gas-phase abundances of all elements, in particular that of oxygen, are solar. As a result, we have found that the ices being formed in the outer Solar Nebula are composed of a mix of clathrate hydrates, hydrates formed above 50 K, and pure condensates produced at temperatures between $\\sim$20 K and $\\sim$50 K. We have noted that, whatever the input parameters adopted in the modelling of the disk, or the formation location considered for icy planetesimals at heliocentric distances beyond 5 AU, their composition remains almost constant, provided that the gas-phase abundances are homogeneous in the nebula. We have argued in this work that gas-drag is responsible for the inward drift of icy particles formed in the outer nebula towards the forming Main Belt. To support this hypothesis, we have showed that, at some epochs of the disk's evolution, some particles produced in the outer nebula may drift to the current position of the Main Belt without encountering temperature and pressure conditions high enough to vaporize the ices they contain. The current distribution of ices potentially existing in asteroids has probably been deeply altered after their formation. The effect of solar insolation may have vaporized the ice within nearer asteroids (semi-major axes of $\\sim$2 AU), melted the ice of mid-range asteroids situated at $\\sim$3 AU, but should not have affected the ice in asteroids located at greater heliocentric distances. Inner and outer asteroids would therefore display no detectable hydratation features, either because the ice was vaporized and dissipated, or because the ice never melted and thus did not react with the surface minerals to a sufficient extent as to allow detection (Cyr et al. 1998). In this context, we have proposed that, from the detection and identification of initially buried ices revealed by recent impacts on the surfaces of asteroids, it could be possible to infer the thermodynamic conditions that occurred within the Solar Nebula during the accretion of these bodies, as well as during the inwards migration of the icy planetesimals which they incorporated. However, this statement requires that either no parent body processing or modification took place during and after the formation of asteroids. For example, we have noted that subsequent alteration of the volatile phases in asteroids may occur due to catalytic reactions in their interiors. We have also investigated the potential influence that the incorporation of ices in asteroids may have on their porosities and densities. In particular, we have showed that the presence of ices can considerably reduce the value of the bulk density of the body, and consequently its macro-porosity, that would be expected from a given taxonomic type. That volatiles were delivered to areas within the ice line is clearly beyond doubt. In addition to the gas-drag mechanism described in this work, it is also likely that a significant amount of volatile material was dynamically driven inwards in the latter stages of planet formation. We still see the tail of this dynamical, chaotic volatile movement today -- the comets we observe passing through the inner solar system are the bearers of ices formed far beyond the snow line, and held in deep freeze since the early days. During the latter stages of planetary migration, the flux of such objects passing through the inner solar system, and hence encountering the asteroids, was significantly higher. Of particular interest, when one considers veneers of volatile material near the surface of the asteroids, is the Late Heavy Bombardment. In the Nice model, (see e.g. Gomes et al, 2005), vast amounts of volatile-rich material is flung inwards from the outer solar system approximately 700 Million years after its birth. This event, caused by the resonant destabilisation of the outer solar system, would have coincided with a simultaneous stirring of the asteroid belt, leading to an impact flux upon the Earth containing approximately even proportions of asteroidal and cometary material. It is clear, though, that the Earth would not be the only object to encounter volatiles injected in this way, and the possibility of a late-veneer of ice arriving in the asteroid belt is surely something which must be acknowledged in future work. In addition to this aggressive and chaotic injection of material, there is also a\u00de gentler mechanism by which volatiles can be driven inwards as a result of planetary migration. As planets migrate, material can be trapped in the locations of mean-motion resonances (MMR), which sweep in front of the planet through it's motion. Evidence of material being swept outwards in the resonances of Neptune is clear for all to see -- the Plutino family of objects are locked in the 2:3 MMR with the planet, and have an inclination distribution which can tell us a great deal about the distance over which the planet migrated, sweeping them along. Inward migration can have the same effect -- the interior resonances of a planet can collect material as it moves inwards, and sweep it along -- giving a mechanism by which volatile material can be eased inwards, with the migration of a giant such as Jupiter. Work such as Fogg \\& Nelson (2006) has shown that such resonant forcing can operate with a resonable efficiency, even for significantly faster migration than expected in our solar system, and so the effects of this behaviour should not be ignored in future work. In spite of the growing pool of evidence pointing towards the existence of water ice in the Main Belt, its detection on asteroids is a challenging observational problem. Large bodies such as Ceres are suspected to have retained a large amount of water since their formation, perhaps even including an internal liquid ocean, throughout the age of the solar system (McCord \\& Sotin 2005). This could particularly be the case if this internal water was originally mixed with some ammonia, in agreement with our composition calculations in section \\ref{icy}, which would have the effect of lowering the melting point of the water (ammonium bearing minerals have been suggested by King et al. (1992) as an alternate explanation for the origin of the 2.07 $\\mu$m band seen in the spectrum of Ceres). Nevertheless, internal water can only be indirectly probed, either by measuring the hydrostatic shape of the object, as was done in the case of Ceres, or by inferring its density from its size and mass, when they are known or, more evidently, from outgasing activities. The case of Ceres is particularly interesting since, in spite of several possible pieces of evidence which support it being a highly hydrated body, the only report of water detection on the dwarf planet is the observation of OH escaping from its northern pole\\footnote{An OH atmosphere was indeed observed around Ceres after perihelion by A'\u00d5Hearn and Feldman (1992) by performing IUE long exposure spectra, with column densities of the order of $10^{11}$~cm$^{-2}$.}, is still not confirmed. Nevertheless, this detection could be explained in the context of the accumulation of ice during winter on the surface or within the subsurface layer, which would then dissipate during summer, when the surface temperature rises. Similar transient events have been suggested as possible mechanisms to trigger the geyser-like activity taking place near the south pole of Enceladus and reported by Cassini (Porco \\& Team 2006). It is interesting to note that, considering the gravity and the day-side temperature of Ceres, any outgassed atmosphere would be rapidly lost. The mean thermal velocity $v_0$ of H$_2$O, for instance, would be close to the escape velocity ($v_{\\infty}=516$~km~s$^{-1}$). Assuming a subsolar temperature of 215~K (Dotto et al. 2000), $v_0$ would vary between 450 to 350~km~s$^{-1}$ from the subsolar point to a zenith angle of 80$^{\\circ}$. As a consequence, hydrodynamical escape would occur ($v_{esc}^{2}/v_{0}^{2} \\le 2$). The photolysis of H$_2$O by solar EUV makes this atmospheric escape even more efficient by giving the photodissociation products OH and H some additional kinetic energy. Considering the short lifetime of H$_2$O at $\\sim$3~AU ($<$9~days), and the fact that the mean thermal velocity of H atoms exceeds $v_{\\infty}$, a tenuous atmosphere of OH is expected if water is outgassed by the asteroid at a sufficient rate. Due to the transcient nature of the atmosphere, the loss of water to space is limited by the flux of water from the interior to the surface. At low latitude, where ice is not stable, the continuous flux of water from the interior to space is too low to be detected. Only an accumulation of water ice at high latitude before perihelion, followed by an outgassing of H$_2$O associated with post-perihelion warming seems to result in an observable column density of OH. These results were found to be consistent with an earlier work done by Fanale and Salvail (1989), who estimated the mean loss rate of H$_2$O to be in the range 30-300~g~s$^{-1}$. Even if one assumes that the atmospheric loss observed by A\u00d5'Hearn and Feldman (1992) occurs continuously at the same rate and at all latitudes (which is obviously wrong as this maximum loss requires high latitudes and post-perihelion conditions), the water loss remains below 4 kg~s$^{-1}$, which, integrated over 4.5~Gyr, corresponds to only 0.07\\% of the mass of Ceres. If the loss rate of H$_2$O in Ceres remained constant throughout its thermal history, the initial water reservoir is thus likely to be integraly preserved. Moreover, since the other volatile species are expected to be trapped as hydrates, clathrate hydrates and pure condensates in this reservoir, we can conclude that they have also been preserved from outgassing throughout the thermal history of the asteroid. Ceres being the largest and, due to its size, probably the wettest Main Belt asteroid, it is an ideal target for carrying out observations aiming at constraining its water regime. The experiment searching for water being vaporised within the polar regions of Ceres should be repeated with the state-of-the-art instrumentation available today on large telescopes. Such a detection would confirm unambiguously the presence of a large amount of water near the surface of Ceres. Direct observation of water ice, or of the effects of hydration, on the surface of Ceres can also be attempted for lower latitudes on the asteroid using a combination of high-angular resolution and spectroscopic instruments permitting the full resolution of its surface to the ~30-40 km level. Due to its low spectral resolution, imaging of the surface of Ceres, even when it is spatially resolved using HST or adaptive optics, is not sensitive to the presence of ice, while the detection of such is within the reach of low resolution spectroscopic observations (e.g. the detection of absorption features in the 1.0-3.5 $\\mu$m region). A spatially resolved spectroscopic mapping of the surface of Ceres in the near-infrared can be done with today's ground-based telescopes and would permit the mapping of the strength of the 3 $\\mu$m band, and allow the search for regions on the surface where interstitial water ice, or hydration features could be present, for instance at the location of cracks within the surface of Ceres, or the locations of deep impact craters. Indeed, recent HST (Thomas et al. 2005) and adaptive optics (Carry et al. 2007) imaging observations of Ceres revealed the presence of large impact craters across its surface which have likely disrupted the outer crust of the asteroid enough to directly expose the sub-surface mantle of wetter material. Finally, a spectroscopic study of the surface of Ceres, in order to search for the spectral signature of water and maybe those of other volatiles, should not be limited to one wavelength region (although the near-infrared range offers many diagnostic bands) but should, instead, encompass a wider range, from the near-UV to infrared wavelengths, in order to improve the identification of the chemicals species responsible for these spectral features. Finally, the NASA Discovery mission {\\it Dawn}, which has been launched in September 2007 and whose arrival at Ceres is scheduled for 2015, will certainly bring new constraints on the presence of volatiles in the Main Belt. In particular, the {\\it Dawn} mapping spectrometer (MS) covers the spectral range from the near UV (0.25 \\micron) through the near IR (5 \\micron) and has moderate to high spectral resolution and imaging capabilities (Russell et al. 2004). These characteristics make it an appropriate instrument for determining the asteroid's global surface composition. Near-infrared mapping of the surface of Ceres at small spatial scales will be very sensitive to volatile concentrations and may reveal ice spots on fresh impact-crater ridges. Moreover, the gravitation investigation of Ceres will allow the determination of its gravity field up to the 12th harmonic degree (Russell et al. 2004). Such a measurement will enable the shape and gravity models to characterize crustal and mantle density variations and, consequently, the amount of volatiles trapped therein." }, "0710/0710.4280_arXiv.txt": { "abstract": "Long Gamma Ray Bursts (GRBs) constitute an important tool to study the Universe near and beyond the epoch of reionization. We delineate here the characteristics of an 'ideal' instrument for the search of GRBs at $z\\ge 6-10$. We find that the detection of these objects requires soft band detectors with a high sensitivity and moderately large FOV. In the light of these results, we compare available and planned GRB missions, deriving conservative predictions on the number of high-$z$ GRBs detectable by these instruments along with the maximum accessible redshift. We show that the {\\it Swift} satellite will be able to detect various GRBs at $z\\ge 6$, and likely at $z\\ge 10$ if the trigger threshold is decreased by a factor of $\\sim 2$. Furthermore, we find that INTEGRAL and GLAST are not the best tool to detect bursts at $z\\ge 6$: the former being limited by the small FOV, and the latter by its hard energy band and relatively low sensitivity. Finally, future missions (SVOM, EDGE, but in particular EXIST) will provide a good sample of GRBs at $z\\ge 6$ in a few years of operation. ", "introduction": "The study of the Universe at the epoch of reionization is one of the main goal of available and future space missions. In the last few years, our knowledge of the early Universe has been enormously increased mainly owing to the observation of Quasars by the SDSS survey (Fan 2006). Long gamma ray bursts (GRB) may constitute a complementary way to study the cosmos and the early evolution of stars avoiding the proximity effects and possibly probing even larger redshifts up to $z\\sim 10$. The five GRBs detected at $z\\gsim 5$, over a sample of about 200 objects observed with the {\\it Swift} satellite (Gehrels et al. 2004), show that a large percentage of GRBs is detected at high-$z$. The current record holder is $z = 6.29$ (Tagliaferri et al. 2005, Kawai et al. 2006). The identification of a large number of GRBs at $z\\ge 6$ will open a new window in the study of the early Universe. Just to give some example, GRBs can be used to constrain the reionization history (Totani et al. 2006, Gallerani et al. 2007), to study the metallicity and dust content of normal galaxies at high-$z$ (Campana et al. 2007), to probe the small-scale power spectrum of density fluctuations (Mesinger, Perna \\& Haiman 2005). Moreover, available and future GRB missions might be the first observatories to detect individual Population III stars, provided that massive metal-free stars were able to trigger GRBs (see Bromm \\& Loeb 2007 for a review). Finally, the study of GRBs at high redshift is interesting by itself. In particular, thanks to cosmological time dilation, the study of the early phases of the afterglow is easier and can provide fundamental highlight on the central engine and burst physics. In this paper, we delineate the main characteristics of an 'ideal' instrument for the search of GRBs at $z\\ge 6-10$. In particular, we explore different observational bands, deriving the best combination of sensitivity and field of view (FOV) in order to detect bursts at $z\\gsim 10$. In the light of these results, we compare available and planned X-- and Gamma--ray missions, deriving conservative predictions on the number of high-$z$ GRBs detectable by these instruments along with the maximum accessible redshift. The paper is organized as follows. In Sect.~2 we briefly describe the different models here adopted. In Sect.~3, we derive the main characteristic of an 'ideal' instrument for exploring the high-$z$ GRB population, whereas predictions for available (planned) GRB missions are given in Sect.~4 (Sect.~5). Finally, we summarize our results in Sect.~6. ", "conclusions": "We have explored the characteristics of an ideal mission for the search of GRBs near and beyond the epoch of reionization, i.e. $z> 6-10$. In particular, we considered different observational bands, deriving the best combination of sensitivity and field of view in order to detect bursts at $z\\gsim 10$. We found that such an experiment requires soft band detectors and high sensitivity, whereas large FOV or wide energy coverage are less important. Assuming 3 sr FOV, an observational band of 8-200 keV, and a sensitivity as low as 0.1 ph s$^{-1}$ cm$^{-2}$, this instrument would be able to detect $\\sim 40$ (3) GRBs at $z\\ge 6$ ($z\\ge 10$) in one year of mission. In the light of these results, we compared available and planned GRB missions, deriving conservative predictions on the observable number of GRBs at $z\\ge 6$ and $z\\ge 10$ along with the maximum accessible redshift. We have shown that {\\it Swift} is a viable tool to detect GRBs at $z\\sim 6$. At the actual trigger threshold, $1.3-4$ GRBs per year should be identified above this redshift. We discuss also the possibility of increasing the number of high-$z$ detections by lowering the {\\it Swift} trigger threshold. We found that the number of detectable GRBs doubles by lowering this by about 50\\%. Assuming to be able to further lowering the trigger threshold down to 0.1 ph s$^{-1}$ cm$^{-2}$, {\\it Swift} should detect $\\sim 1$ GRB per year at $z\\ge 10$. The INTEGRAL and GLAST satellites do not appear the best tools to search for GRBs at very high redshift. The former is limited by the very small FOV whereas the GLAST hard energy observational band and relatively low sensitivity do not allow to detect more than 1 GRB per year at $z\\ge 6$. No GRB at $z\\ge 10$ is expected during the entire mission of both instruments. Finally, we show that future missions, like SVOM, EDGE, and in particular EXIST, will be able to collect a good number of GRBs at $z\\ge 6$ in a few years of operations. This sample can be use to study the early Universe, possible providing strong constrain on the reionization process (Gallerani et al. 2007), and deriving estimate on the star formation and metallicity/dust content in normal high-$z$ galaxies." }, "0710/0710.5529_arXiv.txt": { "abstract": "The strong spectral features near 2.2 $\\mu$m in early-type galaxies remain relatively unexplored. Yet, they open a tightly focused window on the coolest giant stars in these galaxies -- a window that can be used to explore both age and metallicity effects. Here, new measurements of K-band spectral features are presented for eleven early-type galaxies in the nearby Fornax galaxy cluster. Based on these measurements, the following conclusions have been reached: (1) in galaxies with no signatures of a young stellar component, the K-band \\naK\\ index is highly correlated with both the optical metallicity indicator \\mgfe\\/ and the central velocity dispersion $\\sigma$; (2) in the same galaxies, the K-band Fe features saturate in galaxies with $\\sigma > 150$ \\kms\\/ while \\naK\\/ (and \\mgfe) continues to increase; (3) [Si/Fe] (and possibly [Na/Fe]) is larger in all observed Fornax galaxies than in Galactic open clusters with near-solar metallicity; (4) in various near-IR diagnostic diagrams, galaxies with signatures of a young stellar component (strong \\hb, weak \\mgfe) are clearly separated from galaxies with purely old stellar populations; furthermore, this separation is consistent with the presence of an increased number of M-giant stars (most likely to be thermally pulsating AGB stars); (5) the near-IR \\naK\\/ vs.~$\\sigma$ or \\fe\\/ vs.~$\\sigma$ diagrams discussed here seem as efficient for detecting putatively young stellar components in early-type galaxies as the more commonly used age/metallicity diagnostic plots using optical indices (e.g H$\\beta$ vs.~\\mgfe). The combination of these spectral indices near 2.2 $\\mu$m with high {\\it spatial} resolution spectroscopy from ground-based or space-based observatories promises to provide new insights into the nature of stellar populations in the central regions of distant early-type galaxies. ", "introduction": "\\label{sec:intro} Understanding the stellar content of early-type galaxies is fundamental to understanding their star formation and chemical evolution history. Most early-type galaxies are too distant to resolve their individual stars with current technology, rendering the direct study of their stellar populations impossible. Thus, their stellar populations must be studied using indirect methods. In recent decades, significant effort has gone into trying to better constrain the stellar contents for early-type galaxies using optical spectroscopic data. The most commonly studied features have been Ca I H and K 0.38 $\\mu$m, H$\\beta$, Mgb 0.52 $\\mu$m, Fe $\\mu$m 0.53, Na 0.82 $\\mu$m, and CaT 0.86 $\\mu$m. Interpretation of all such spectral features is intrinsically complicated by their blended nature -- each feature is really the super-position of many spectral lines, usually from several different elements, blurred together by the line-of-sight velocity dispersion within each galaxy. There is no way to overcome this problem -- it must simply be taken into account during analysis. As population synthesis models have become more sophisticated and digital stellar libraries more complete, this problem has become more tractable over time. Another challenge arises from the composite nature of galaxies: each observed feature is the luminosity-weighted integrated sum of that feature from all stars in the observed line-of-sight. Naturally, luminosity-weighted does not imply mass-weighted. A relatively small fraction of the mass can dominate the observed luminosity and mask the underlying stellar population (e.g. as happens during a starburst event within a pre-existing galaxy). Even in relatively quiescent galaxies, light from stars at several important evolutionary stages contribute roughly equally to the observed spectral features between 0.4 -- 1 $\\mu$m range. Hence, a feature depth change could be due to (e.g.) a change near the (mostly) age-driven main-sequence turnoff or the (mostly) metallicity-driven red giant branch. The details can become quite complicated, as illustrated by the long standing controversy about whether observed changes in Balmer line strength arise from the presence of younger main sequence stars, more metal-poor main sequence stars, or an extended horizontal giant branch (for recent discussions of this debate, see Maraston \\& Thomas 2000 and Trager et al. 2005). A similar controversy surrounds Na 0.82 $\\mu$m feature: is it driven by metallicity-driven red giant branch changes, initial mass function related differences in the relative number of cool dwarf and giant stars or both \\citep[e.g.][]{car86, all89, del92}? However, the properties of the RGB component can be isolated by observing in the K-band (centered near 2.2 $\\mu$m). At those wavelengths, cool giants near the tip of the first-ascent red giant branch (RGB) dominate the integrated light in old ($\\geq$ 3 Gyr) stellar populations. In combination with optical observations, K-band observations should facilitate the separation of MSTO and RGB light contributions. There are two possible complications to this scenario. First, a very young stellar population containing red supergiants will contribute a significant fraction K-band light. Fortunately, such a population is obvious from the presence of H~II region emission lines at shorter wavelengths. Second, a somewhat older population (1 -- 2 Gyr, i.e. an intermediate-age population) may contain bolometrically bright carbon stars that can contribute a detectable amount of K-band light (see discussions in Silva \\& Bothun 1998a,b). Such a population may or may not be connected to increased H$\\beta$ strength. Initial development of these ideas can be found in Silva et al. (1994), Mobasher \\& James (1996), James \\& Mobasher (1999), Mobasher \\& James (2000), all of whom focused on the CO 2.36 $\\mu$m feature. Origlia et al. (1997) observed a Si dominated feature at 1.59 $\\mu$m as well as CO dominated features. These observational studies were limited by small-format detectors to relatively low resolving powers and/or small wavelength ranges per observation. In the cases of Silva et al. and Origlia et al., only small, heterogeneous samples of galaxies were observed. A general conclusion of the James \\& Mobasher studies was that changes in CO strength between early-type galaxies in high-density and low-density regions were statistically consistent with different fraction contributions of intermediate-age AGB light and hence galaxies in low-density regions had younger luminosity-weighted mean ages. Origilia et al. argued that [Si/Fe] was super-solar in the four elliptical galaxies they observed. To further develop these ideas and investigate the usefulness of other K-band spectral indices in the study of early-type galaxies, new data have been obtained for eleven E/S0 galaxies in the nearby Fornax cluster. Only measurements in the central regions of these galaxies are discussed here. In Section~\\ref{sec:data}, the galaxy sample and its observations are discussed, while in Section~\\ref{sec:proc} the data processing methodology is described. The measurement of spectral feature strength is explained in Section~\\ref{sec:lines} while basic observation results are presented in Section~\\ref{sec:results}. The broader astrophysical implications of our observational results are discussed in Section~\\ref{sec:disc}. A summary is provided at the end. ", "conclusions": "\\label{sec:disc} No self-consistent theoretical spectral synthesis models for the interpretation of integrated K-band spectra of early-type galaxies are widely available. Development of such models is on-going (e.g. Marmol Queralt, 2007, in preparation). Until such models exist, we must rely on basic (and perhaps imperfect) astrophysical intuition and available tools, such as the high-resolution near-IR spectral atlas of Wallace \\& Hinkle (1996) (hereafter WK96). \\subsection{Galaxies with young populations} Three galaxies in our sample (NGC 1316, NGC 1344, and NGC 1375) have stronger \\hb\\/ and weaker \\mgfe\\/ features than the rest of our sample (see Figure~\\ref{fig:fornax_pops}). This suggests a simple, two component stellar population model. One stellar component is cold -- in the luminosity-weighted mean, it is presumably old ($>$ 8 Gyr) and metal-rich ([Fe/H] $>$ --0.3) with spectral properties consistent with the observed central velocity dispersion. In other words, it has properties similar to the purely old galaxies in our sample (see Figure~\\ref{fig:fornax_pops}). The other component is warm -- it contains a significant number of A/F dwarf or sub-giant stars. This component could have several origins but only two are considered here. Each has different consequences for K-band spectral index behavior. First, the warm component could be associated with a young population with a warm ($\\sim$2 \\msun) main sequence turnoff (MSTO). The MSTO stars are tied to thermally pulsating asymptotic giant branch (TP-AGB) stars with bolometric magnitudes that place them above the tip of the first-ascent red giant branch (TRGB). The TP-AGB stars are cooler than stars on the first-ascent RGB and hence have an M spectral type. In the underlying luminosity function (spectral type vs. number of stars), the M-star bins are relatively more populated than in a purely old galaxy. The net effect is that integrated 2.2 $\\mu$m spectra should become more M-like. Second, the warm component could be associated with a metal-poor population with a warm MSTO. These MSTO stars are tied to RGB stars with similar luminosity as the corresponding metal-rich RGB stars but with warmer \\teff. In the underlying luminosity function, the K-star bins will be relatively more populated than in a purely old population. The net effect is to make the integrated 2.2 $\\mu$m spectra more K-like. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=0]{f16.ps}} \\caption{\\label{fig:galModel} Optical-NIR corrections for young populations. See Figure~\\ref{fig:galLineSigma} for a symbol explanation. The observed index strengths for the galaxies with young components are connected to the predicted index strengths ({\\it solid squares}) after the effects of a young component has been removed. In the lower panel, the {\\it slanted dashed line} indicates the unweighted linear regression fit for the purely old galaxies with $\\sigma >$ 150 \\kms. In the middle panels, the {\\it horizontal dashed line} indicates the unweighted mean value for purely old galaxies with $\\sigma >$ 150 \\kms.} \\end{figure} Can the observations presented here distinguish between these two scenarios? Consider Figure~\\ref{fig:galModel} -- the galaxies with relatively strong \\hb\\/ and weak \\mgfe\\/ have relatively stronger (i.e. more M-like) near-IR features. Qualitatively, this is consistent with the first scenario. Can this conclusion be better quantified? What is the relative ratio of young to old stellar mass? Are the changes in optical and near-IR spectral features {\\it quantitatively} consistent? To begin to answer these questions, the Thomas et al. (2003) models were used to create very simple two-component models that produced optical index values that matched the observed values in the three galaxies with strong \\hb. One component was old (11 Gyr) with metallicity set to [Z/H] $=$ 0.35 for NGC 1316 and NGC 1344 and 0.0 for NGC 1375. The other component had the same metallicity but a young (1 Gyr) age. The relative mass fractions of these components were varied until the model \\hb\\/ strength matched the observed \\hb\\/ strength. The resultant young mass fractions (where total mass $=$ 1) were 0.135, 0.075, and 0.140 for NGC 1316, NGC 1344, and NGC 1375, respectively. The implied differential correction between the purely old model and the two-component model was then applied to the observed data (see Figure~\\ref{fig:galModel}). Next, observed \\fe\\/ strength was adjusted manually until all three galaxies lay within the locus of purely old galaxies with similar $\\sigma$ in the \\fe\\/ vs. \\naK\\/ panel. As part of this adjustment, $\\Delta$(\\fe)/$\\Delta$(\\naK) was forced to agree with the value (0.62) determined for the observed Galactic cluster giant stars (see Figure~\\ref{fig:starLineLine}). This is a trend in effective temperature that is relatively insensitive to metallicity. Cooler giant stars (more late K and early M like) have strong K-band spectral features. In addition to the $\\Delta$(\\fe)/$\\Delta$(\\naK) measured from the Galactic open cluster stars, two other key numerical trends can be computed: \\begin{eqnarray} \\Delta(\\fe)/\\Delta(\\naK) & = & 0.62 \\\\ \\Delta(\\naK)/\\Delta(\\mgfe) & = & 0.70 (1.10) \\\\ \\Delta(\\fe)/\\Delta(\\mgfe) & = & 0.28 (0.44) \\end{eqnarray} NGC 1316 and NGC 1344 could be forced to have consistent trajectories. However, NGC 1375 (values shown in parenthesis) appears to follow somewhat different trajectories. Obviously, this kind of cartoon model is illustrative only and surely hides a plethora of details. Indeed, the real trajectories are unlikely to be linear. Nevertheless, a clear astrophysical conclusion emerges: for the \\hb-strong galaxies, the observed optical and near-IR features all change in concert, consistent with a warm MSTO tied to an extended RGB, which is in turn consistent with the presence of a young stellar component, not a metal-poor stellar component. \\subsection{Galaxies with bright K-band SBF} Liu, Graham, \\& Charlot (2002) found that relative I-band and K-band surface brightness fluctuation strength (expressed as an SBF color, $\\bar{I} - \\bar{K}$) was essentially constant for the majority of the early-type galaxies they studied in Fornax. However, two galaxies in common with our sample (NGC 1419 and 1427) were found to have brighter (larger) K-band SBF relative to the mean relationship (see also Mei et al. 2001). A larger $\\bar{K}$ is thought theoretically to arise from the presence of an extended giant branch, i.e. more cool, bright, M giants. In the spectroscopic data discussed here, there are no indications of such extended giant branches in these galaxies -- their central line indices are consistent with an integrated stellar populations dominated by an old, metal-rich population. Of course, the SBF measurements were made in the outer regions of these galaxies, as opposed to the optical and near-IR observations of the central regions discussed here. It may be that the integrated luminosity of such a component (if present at all) is not high enough in the central region for detection by our method. \\subsection{Purely old galaxies} In purely old galaxies, we have seen that: \\begin{itemize} \\item{} \\naK\\ is stronger than in Galactic open clusters and is highly correlated with $\\sigma$ and \\mgfe. \\item{} \\fe\\/ saturates for $\\sigma \\gtrsim 150$ \\kms. \\item{} \\co\\ is somewhat correlated with $\\sigma$ and \\mgfe, while \\caK\\/ is not (unless NGC 1399 is not included in the regression fits, see footnote 2 above). \\end{itemize} The observed K-band spectral features are named for the dominant elemental species in the {\\it solar} spectrum. However, at the effective temperatures of interest here, the on-band index definition and their companion off-band continuum bands contain lines from other elemental species as well (as Ramirez et al.~1997 discuss comprehensively). By referring to the high resolution spectra of WK96, the contribution from these other absorbers can be investigated and used to explain the overall spectral feature behavior. In turn, underlying astrophysical parameters related to galaxy formation and evolution are revealed. \\subsubsection{\\naK\\/ index} In the current galaxy sample, \\naK\\/ is significantly stronger than observed in the Galactic open cluster stars. The off-bands defined for the \\naK\\/ feature are relatively line-free, so we can focus on the on-band spectral region (see the $\\lambda$ Dra, M0 III spectrum in WK96, p. 352). In addition to sodium, Sc, Si, and (to a lesser extend) V are significant absorbers in this region. Origlia et al. (1997) argued that [Si/Fe] is super-solar in a small sample of early-type galaxies based in measurements of the Si I 1.59 $\\mu$m feature. Trager et al. (2000) have argued that both [Na/Fe] and [Si/Fe] are super-solar in early-type galaxies based on models of optical spectral feature behavior. The K-band \\naK\\/ measurements presented here are consistent with enhanced silicon in the observed Fornax galaxies and hence indirectly confirm the conclusions of Origlia et al. and Trager et al. \\subsubsection{\\fe\\/ index} The \\fea\\/ feature contains approximately equal absorption contributions from Fe, Sc and Ti while \\feb\\/ is dominated by Fe absorption features with some additional Sc absorption (in WK96, see the $\\lambda$ Dra, M0 III spectra on p. 346 and 348). Hence, the \\fe\\/ index (a combination of \\fea\\/ and \\feb) is dominated by absorption lines from Fe-peak elements). For purely old galaxies with $\\sigma >$ 150 \\kms, \\fe\\/ is observed to remain constant while \\naK\\/ (an index dominated by absoprtion lines from $\\alpha$-elements) and \\mgfe\\/ (a total metallicity [Z/H] indicator) become stronger as $\\sigma$ increases. It is tempting to conclude that above $\\sigma >$ 150 \\kms\\/ luminosity-weighted mean total metallicity [Z/H] continues to increase as a function of central velocity dispersion, driven by a relative increase of [$\\alpha$/H], while [Fe/H] remains constant. However, within the narrow [Fe/H] range of observed Galactic open cluster stars, \\fe\\/ is more strongly correlated with effective temperature than cluster [Fe/H] (see Figure~\\ref{fig:starLineColor}). Increasing mean total metallicity [Z/H] in the galaxies should correspond to cooler mean RGB effective temperature and hence increased \\fe. Recall that the relationship between \\fe\\ and mean RGB effective temperature has already been exploited above to explain the observed behavior of \\fe\\/ (and \\naK) in galaxies with young stellar components. The dilemma is clear: how can a temperature-sensitive index like \\fe\\/ remain constant while mean RGB effective temperature decreases due to increasing [Z/H]? Without appropriate population synthesis models or stellar spectra, only a few speculative thoughts can be offered. In their detailed comparison of non-solar aboundance population synthesis models with measurements of spectral indices in the optical Lick system, Trager et al. (2000) concluded that their best-fit models included enhancements in C, N, Na, and Si (among others). As [$\\alpha$/H] increases with $\\sigma$ (as suggested by increasing \\naK), stronger CN bands in the \\fea\\/ and \\feb\\/ on-band and off-band windows (see the WK96 spectra referenced above and the band definitions in Table~\\ref{tab:features}) may have the effect of depressing the local continuum and hence decreasing the value of \\fe\\/ $=$ (\\fea\\/ $+$ \\feb)/2. For example, Trager et al. noted a similar effect from C$_{\\rm 2}$ bands in the optical Mg {\\it b} index. Therefore, it may be that above some critical mean [$\\alpha$/H] and below some critical mean effective temperature, \\fe\\/ stays within a narrow range and provides no useful information about mean [Z/H], [Fe/H], or effective temperature. A conclusive astrophysical explanation of \\fe\\/ in early-type galaxies with $\\sigma >$ 150 \\kms\\/ awaits larger galaxy samples, observations of appropriate stars (e.g. in the Galactic bulge), and detailed population synthesis models. \\subsubsection{\\co\\ index} The \\co\\/ feature is dominated by the $^{\\rm 12}$CO(2,0) bandhead and has no significant absorption components from other elements. Like \\fe\\/ above, \\co\\/ is highly correlated with $T_{eff}$ in the observed Galactic open cluster stars and reaches larger values than observed in the galaxies. As a function of central velocity dispersion $\\sigma$ in the observed galaxies, \\co\\/ does not appear to saturate (or at least not as definitively as \\fe). This adds credence to the suspicion that \\fe\\ is not tracing mean RGB effective temperature in galaxies with $\\sigma >$ 150 \\kms. Based on the current measurements, no conclusions can be reached about relative carbon or oxygen abundance in these galaxies. \\subsubsection{\\caK\\ index} No obvious explanation presents itself for the behavior of \\caK\\/ in this galaxy sample. The \\caK\\/ feature is intrinsically complex, consisting of contributions from many absorbers: Ca, S, Si, and Ti (all $\\alpha$-elements) as well as Sc and Fe (see $\\lambda$ Dra, MO III spectrum in WK96, p. 342). As a function of \\teff\\/ in the range of interest, some of the contributing absorption lines get stronger (Sc and Ti, the former faster than the latter), some features get weaker (Si), and some remain roughly constant (Ca, S, and Fe) (cf. Ramirez et al. 1997). Relative to solar abundance ratios, Trager et al. (2000) have argued that Si and S are over-abundant and Ca is under-abundant in early-type galaxies while [Ti/Fe] and [Sc/Fe] have their solar values. In the Galactic open cluster stars observed here, the \\caK\\/ feature is stronger at a given $J-K$ ($T_eff$) in the solar-metallicity stars than in the more metal-poor stars. Within the galaxies observed here, \\caK\\/ appears to have significant scatter at any given $\\sigma$ or \\mgfe. No obvious explanation for this scatter (or the global behavior of \\caK\\ with $\\sigma$ and \\mgfe) presents itself at this time. Using new, moderate resolution ($R \\sim 2500$) K-band spectra, spectral indices have been measured in the central regions of eleven early-type galaxies in the nearby Fornax cluster. Based on these measurements, the following conclusions were reached: \\begin{enumerate} \\item{} The \\naK\\/ feature is much stronger in these Fornax early-type galaxies than observed in solar-metallicity Galactic open cluster stars. This is attributed to relative [Si/Fe] (and possible [Na/Fe]) differences between the open cluster stars and the Fornax galaxies, i.e. both are larger in the Fornax galaxies than in the cluster stars. \\item{} In various near-IR diagnostic diagrams, galaxies with optical indices indicative of a warm stellar component are clearly separated from galaxies dominated by colder, presumably old ($\\geq$ 8 Gyr) stellar populations. Changes in the near-IR spectra features are consistent with the presence of an cool component dominated by late K and/or early M giants stars. In combination, the optical and near-IR observations are consistent with the presence of a young stellar component with a warm MSTO and a significant extended giant branch consisting of TP-AGB stars. \\item{} For detecting a young stellar component, the \\naK\\/ vs.~$\\sigma$ or \\fe\\/ vs.~$\\sigma$ diagnostic diagram seems as efficient as using H$\\beta$ vs. \\mgfe\\/ (or other similar combinations of optical indices). The near-IR features have the additional advantage that no emission-line correction is needed (as it is necessary for the stronger optical Balmer lines). \\item{} The \\fe\\/ index saturates in galaxies with central velocity dispersion $\\sigma$ $>$ 150 \\kms\\/ dominated by old ($\\geq$ 8 Gyr) stellar populations. For $\\sigma >$ 150 \\kms, these Fe features are unlikely to be useful for investigating stellar population differences between early-type galaxies. Above $\\sigma >$ 150 \\kms, the continued increase in \\naK\\ (and \\mgfe) strength with $\\sigma$ coupled with constant \\fe\\ presents an astrophysical challenge. Although it is tempting to conclude that [Fe/H] reaches a maximum value, while [$\\alpha$/H] (and hence [Z/H]) continues to increase, more observational and population synthesis work is needed to understand conclusively the behavior of \\fe\\/ in these high-mass early-type galaxies. \\end{enumerate} Adding these near-IR indices to the standard diagnostic toolkit for analyzing the integrated light of early-type galaxies clearly has great potential. To develop this potential, three obvious steps are needed: observe more galaxies over a larger range of central velocity dispersion (including field galaxies with existing optical data) and extend current population synthesis models to the near-IR. We know that various groups are working on both steps. It would also be useful to study the radial behavior of these indices within individual galaxies to compare and contrast index behavior between galaxies. For the foreseeable future, the study of the central populations in early-type galaxies will remain the study of integrated light. As distance increases, our ability to study the central regions of early-type galaxies within metric apertures equivalent to nearby galaxies relies on the high {\\it spatial} resolution spectroscopy achievable from space or with adaptive optics on the ground. In the former case, the James Webb Space Telescope represents the frontier -- yet, no optical spectrograph that works below 0.8 $\\mu$m is currently planned. In the latter case, the frontier will be shaped by further development of adaptive optics systems -- and these systems will achieve their best performance beyond 1 $\\mu$m. Hence, understanding how to interpret near-IR spectral indices in nearby galaxies is key to facilitating the kind of investigation and characterization of more distant early-type galaxies that will not be possible at optical wavelengths." }, "0710/0710.4059.txt": { "abstract": "The turbulent magnetic diffusivity tensor is determined in the presence of rotation or shear. The question is addressed whether dynamo action from the shear--current effect can explain large-scale magnetic field generation found in simulations with shear. For this purpose a set of evolution equations for the response to imposed test fields is solved with turbulent and mean motions calculated from the momentum and continuity equations. The corresponding results for the electromotive force are used to calculate turbulent transport coefficients. The diagonal components of the turbulent magnetic diffusivity tensor are found to be very close together, but their values increase slightly with increasing shear and decrease with increasing rotation rate. In the presence of shear, the sign of the two off-diagonal components of the turbulent magnetic diffusion tensor is the same and opposite to the sign of the shear. This implies that dynamo action from the shear--current effect is impossible, except perhaps for high magnetic Reynolds numbers. However, even though there is no alpha effect on the average, the components of the $\\alpha$ tensor display Gaussian fluctuations around zero. These fluctuations are strong enough to drive an incoherent alpha--shear dynamo. The incoherent shear--current effect, on the other hand, is found to be subdominant. ", "introduction": "Many of the stellar and planetary magnetic fields are believed to be the result of a dynamo process that converts kinetic energy from turbulent motions and shear into magnetic energy. A particular challenge consists in explaining the field on length scales that exceed the scale of the turbulence. This topic has traditionally been addressed within the framework of mean--field electrodynamics (Krause \\& R\\\"adler 1980). Over the decades the applicability of this theory has repeatedly been questioned (e.g., Piddington 1981, Vainshtein \\& Cattaneo 1992). Meanwhile, direct simulations of hydromagnetic turbulence have begun to show dynamo action (Meneguzzi et al.\\ 1981, Meneguzzi \\& Pouquet 1989, Nordlund et al.\\ 1992, Brandenburg et al.\\ 1996, Cattaneo 1999). In some particular cases, large-scale fields are being generated (Glatzmaier \\& Roberts 1995, Brandenburg et al.\\ 1995, Brandenburg 2001) which raises the question about the mechanism responsible for this phenomenon. In cases where the flow is systematically non-mirror symmetric the association with an $\\alpha$ effect is obvious. However, there are now also examples of nonhelical large-scale dynamos owing to turbulence under the influence of shear alone (Brandenburg 2005a, Yousef et al.\\ 2007). Their interpretation is not straightforward, because several possible mechanisms have been proposed that might produce dynamo action from turbulence and shear alone, i.e.\\ {\\it without} rotation and stratification that otherwise would have been the main ingredients of an $\\alpha$ effect. The most detailed investigations have been carried out in connection with the so-called shear--current effect (Rogachevskii \\& Kleeorin 2003, 2004, R\\\"adler \\& Stepanov 2006, R\\\"udiger \\& Kitchatinov 2006). Another possibility is a magnetic $\\alpha$ effect that is driven by a current helicity flux, as was suggested by Vishniac \\& Cho (2001; see also Brandenburg \\& Subramanian 2005c). A third possibility might be an incoherent (random) $\\alpha$ effect with zero mean and finite variance, suggested by Vishniac \\& Brandenburg (1997) in connection with accretion discs (see also Sokolov 1997, Silant'ev 2000, Fedotov et al.\\ 2006, Proctor 2007). The only reliable way to determine what is the dominant effect is to calculate all relevant components of the $\\alpha$ and turbulent magnetic diffusivity tensors in a general expansion of the electromotive force in terms of the mean magnetic field. The case considered in Brandenburg (2005a) is unnecessarily complicated because the shear employed there depends on two Cartesian coordinates. A simpler possibility is to consider a shear flow depending linearly on only one coordinate and we shall pursue this idea in the present paper. The shear--current effect and the incoherent $\\alpha$ effect could then still operate. Because we will use periodic boundary conditions there can be no magnetic helicity flux, so the Vishniac \\& Cho (2001) effect is then ruled out, even though it could still, at least in principle, explain the generation of a mean magnetic field in the simulations of Brandenburg (2005a), which do possess a helicity flux. In this paper we calculate all relevant components of $\\alpha_{ij}$ and $\\eta_{ijk}$ using the so-called test field method. This method was introduced by Schrinner et al.\\ (2005, 2007) in connection with convection in a spherical shell and used later by Brandenburg (2005b), Sur et al.\\ (2007) and Brandenburg et al.\\ (2008) in connection with forced turbulence in Cartesian boxes. The essence of this method consists in solving evolution equations for the fluctuations of the magnetic field around suitably defined test fields such that all relevant coefficients can be computed. ", "conclusions": "The present work has demonstrated that the test field method provides a robust means of determining all components of the turbulent magnetic diffusivity tensor that are relevant for mean fields depending only on $z$ and $t$. Both rotating and weakly shearing turbulence are studied. In either case the diagonal components of the turbulent diffusivity tensor are about equal to each other. Shear slightly enhances the turbulent magnetic diffusivity while rotation quenches it. In the presence of rotation, the $\\OO\\times\\meanJJ$ effect occurs, which is described by the off-diagonal components of the turbulent magnetic diffusivity tensor. Shear leads to the shear--current effect, again described by off-diagonal components of this tensor. In both cases the results are consistent with those found in the framework of the second-order correlation approximation. The possibility of the so-called shear-current dynamo has been scrutinized. It depends crucially on the sign of the component $\\eta_{21}$ of the magnetic diffusivity tensor. It turns out that, within the ranges of parameters considered, its sign is in general not suited for driving a dynamo based on this effect, with a possible exception at large magnetic Reynolds numbers. In this way the analytic results found in the second--order correlation approximation for incompressible fluids (R\\\"udiger \\& Kitchatinov 2006; R\\\"adler \\& Stepanov 2006) are confirmed and generalized. Direct numerical simulations are presented which exhibit growing mean magnetic fields in shear flow turbulence. An interpretation as a (coherent) shear-current dynamo is hardly possible. Instead, it is argued that it can be explained by an incoherent alpha--shear dynamo. The incoherent shear--current effect has also been determined, but it is found to be less important." }, "0710/0710.0760_arXiv.txt": { "abstract": "{ This series of papers is dedicated to a new technique to select galaxies that can act as standard rods and standard candles in order to perform geometrical tests on large samples of high redshift galaxies to constrain different cosmological parameter. The goals of this paper are (1) to compare different rotation indicators in order to understand the relation between rotation velocities extracted from observations of the H${\\alpha} \\lambda$6563\\AA\\ line and the [OII]$\\lambda 3727$\\AA\\ line, and (2) to determine the scaling relations between physical size, surface brightness and magnitude of galaxies and their rotation velocity using the SFI++, a large catalog of nearby galaxies observed at I-band. A good correlation is observed between the rotation curve-derived velocities of the H${\\alpha}$ and [OII] observations, as well as between those calculated from velocity histograms, justifying the direct comparison of velocities measured from H${\\alpha}$ rotation curves in nearby galaxies and from [OII] line widths at higher redshifts. To provide calibration for the geometrical tests, we give expressions for the different scaling relations between properties of galaxies (size, surface brightness, magnitude) and their rotation speeds. Apart from the Tully-Fisher relation, we derive the size-rotation velocity and surface brightness-rotation velocity relations with unprecedentedly small scatters. We show how the best size-rotation velocity relation is derived when size is estimated not from disc scale lengths but from the isophotal diameter $r_{23.5}$, once these have been corrected for inclination and extinction effects.} \\titlerunning{Geometrical Tests of Cosmological Models II.} \\authorrunning{Saintonge et al.} ", "introduction": "With the new generation of telescopes and instruments, such as multi-object spectrograph and large field of view cameras, it has become possible to perform large, deep redshift surveys. Using these data, the structure and geometry of the Universe can be studied through new channels, namely geometrical tests such as the angular diameter test (angular size - redshift relation), the Hubble diagram (magnitude - redshift relation) and the Hubble test (count - redshift relation). These tests are generally difficult, requiring differentiation between the natural evolution of galaxies and the effects of geometry. Large redshift surveys ease the task in that respect. To perform the geometrical tests mentioned above, a population of objects that can be tracked through redshift needs to be defined. More specifically, the angular diameter test requires a standard rod to be identified, the angular size of which is then being measured over the redshift range of interest. The objects taken to serve that purpose could be anything from galaxies \\citep[e.g.][]{sandage72}, clusters of galaxies \\citep[e.g.][]{sereno,cooray98}, or dark matter halos \\citep[]{cooray01}. Regardless of its nature, a standard rod should satisfy simple criteria: its size should be measured reliably up to high redshifts, and it should be observable in the local Universe to provide calibration. In this study, the relation between the physical size of a spiral galaxy and its rotation velocity (or global profile width) is used. By virtue of this relation, selecting a population of objects with a given rotation velocity is equivalent to selecting them based on their physical size. In other papers of this series \\citep[][thereafter Paper I and Paper III]{paperI,paperIII}, the angular diameter test is performed on a dataset of precursor Vimos/VLT Deep Survey (VVDS) \\citep[]{vvds} galaxies, using the linewidth-diameter relation to select standard rods. The test is used to isolate the effects of disk evolution with redshift, and to constrain the values of cosmological parameters. Though the VVDS provides excellent data at high redshift, the angular diameter test highlights one of its shortcomings: limited volume coverage at low redshift and therefore the lack of a comparison sample in the local Universe. In this context, the goal of this paper is twofold. First, using a large sample of nearby galaxies \\citep[SFI++;][]{spr07} the physical properties of the standard rods and standard candles used to perform the geometrical tests are established, free from any evolutionary biases. This is especially useful to calibrating the angular diameter relation and to separating the effects of galaxy evolution from those of geometrical variations (see Paper I). By establishing clearly the scaling relations between the size, surface brightness and magnitude of these nearby galaxies observed at I-band and their rotation velocity, we establish the calibration that will allow us to select standard rods/candles simply by tracing through redshift galaxies with a given rotation speed. Having a strong handle on the structural parameters of disc galaxies in the local universe gives the unique opportunity of tracing over time the evolution of these properties for disc galaxies hosted in dark matter halos of the same mass, if rotation velocity is used as a proxy for halo mass, and compare them with predictions made under the hierarchical scenario for the growth of structures \\citep[e.g.][]{mo98}. The second goal of this paper is to compare the rotation velocity indicator used for the high redshift data to those used in the local Universe, cross-calibrating rotation indicators used at different redshifts. This last point is of relevance since rotation information for spiral galaxies can be obtained through the observation of various spectral lines, the choice of which varies with the redshift of the sample. For galaxies with $z \\lesssim 0.1$, the HI 21 cm line is an excellent candidate. The H${\\alpha} \\lambda$6563\\AA\\ line is also frequently used for galaxies with low to moderate redshifts. However, the H${\\alpha}$ line is quickly redshifted into the near-infrared and becomes unavailable to ground observers using optical telescoptes for galaxies with $z \\gtrsim 0.4$, even though with new near-infrared spectrographs such as SINFONI it is now possible to obtain H$\\alpha$ rotation curves for high redshift galaxies \\citep[e.g.][]{forster06}. Even with the advent of such instruments, most large studies of galaxies at high redshift rely on the [OII]$\\lambda 3727$ \\AA\\ line, including the VVDS and the DEEP2 Redshift Survey \\citep[]{davis01}. In order to compare sets of local and distant galaxies, it is therefore necessary to understand how rotation velocities extracted from these different lines relate. It has already been shown that velocity widths derived from HI 21cm and H${\\alpha}$ observations are in excellent agreement \\citep[see for example][]{courteau97,vogt}. The correlation between the [OII] and HI line widths has also been investigated \\citep[]{kobulnicky}. Using a sample of 22 nearby late-type spirals they find [OII] widths to be accurate to within $10 \\%$ for galaxies with a roation width of 200 km s$^{-1}$, which is comparable to the overall scatter in the local Tully-Fisher relation. However, they conclude that the uncertainties go up to about 50\\% for galaxies with widths $<150$ km s$^{-1}$. Here, a sample of 32 spiral galaxies with $0.155 1$ is achieved for a DTD with a Gaussian distribution centered at about 3~Gyr. Nevertheless this DTD fails to reproduce the dependence of the SN rate on galaxy colours which is observed in the local Universe \\citep{Mannucci06}. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{delay.eps}} \\caption{Delay time distributions as derived by \\citet{Greggio05} for SD and DD models, \\citet{Mannucci06} and \\citet{Strolger}.} \\label{delay} \\end{figure} The selected DTD functions are plotted in Fig.~\\ref{delay} while the predicted evolutionary behaviours of the SN~Ia rate are compared with all published measurements in Fig.~\\ref{rateia}. In all cases, the value of $k_{\\alpha}A^{\\rm Ia}$ was fixed to match the value of the local rate; depending on the model it ranges between 3.4-7.6$\\times 10^{-4}$. This normalization implies that, for the adopted SalA IMF, and assuming a mass range for the progenitors of $3-8$ $M_\\odot$, the probability that a star with suitable mass becomes a SN~Ia, is $\\sim$ 0.01-0.03. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{rate_ia.eps}} \\caption{SN~Ia rate measurements fitted with different DTD functions and the SFH by \\citet{Hopkins06}. Symbols for measurements are as in Fig.~\\ref{obsrate}.} \\label{rateia} \\end{figure} The models obtained with the different DTDs are all consistent with the observations with the exception of the \"wide\" DD model, whose redshift evolution is definitely too flat. On the other hand none of the DTD functions, with the adopted SFH, is able to reproduce at the same time the very rapid increase from redshift 0 to 0.5 suggested by some measurements \\citep{Barris,Dahlen04} and the decline at redshift $>1$ \\citep{Dahlen04}. We note that a new measurement of \\citet{Poznanski} suggests that the SN~Ia rate decline at high redshift may be not as steep as estimated by \\citet{Dahlen04}. Given the current uncertainties of both SN~Ia rate and SFH it is difficult to discriminate between the different DTD functions and hence between the different SN~Ia progenitor models. To improve on this point, more measurements of SN~Ia rate at high redshifts are required to better trace the rate evolution. At the same time measurements in star forming and in passive evolving galaxies in a wide redshift range can provide important evidence about the SN~Ia progenitor models. In addition, it is essential to estimate the cosmic SFH more accurately because the position of the peak of the SFH was found to be the crucial parameter for the recovered delay time \\citep{Forster}." }, "0710/0710.1550_arXiv.txt": { "abstract": "We present VLT/FORS2 spectroscopic observations of globular clusters (GCs) in five low surface brightness (LSB) dwarf galaxies: KK211 and KK221, which are both dwarf spheroidal satellites (dSph) of NGC~5128, dSph KK84 located close to the isolated S0 galaxy NGC~3115, and two isolated dwarf irregular (dIrr) galaxies UGC~3755 and ESO~490-17.~Our sample is selected from the Sharina et al. (2005) database of Hubble Space Telescope WFPC2 photometry of GC candidates in dwarf galaxies. For objects with accurate radial velocity measurements we confirm 26 as genuine GCs out of the 27 selected candidates from our WFPC2 survey.~One candidate appears to be a distant galaxy.~Our measurements of the Lick absorption line indices in the spectra of confirmed GCs and the subsequent comparison with SSP model predictions show that all confirmed GCs in dSphs are old, except GC KK211-3-149 ($6 \\pm$2 Gyr), which we consider to be the nucleus of KK211. GCs in UGC~3755 and ESO~490-17 show a large spread in ages ranging from old objects ($t>10$ Gyr) to clusters with ages around 1 Gyr. Most of our sample GCs have low metallicities $\\zh \\le -1$. Two relatively metal-rich clusters with $\\zh \\approx -0.3$ are likely to be associated with NGC~3115. Our sample GCs show in general a complex distribution of $\\alpha$-element enhancement with a mean $\\langle$[$\\alpha$/Fe]$\\rangle=0.19\\pm0.04$ derived with the $\\chi^{2}$ minimization technique and $0.18\\pm0.12$ dex computed with the iterative approach. These values are slightly lower than the mean $\\langle$[$\\alpha$/Fe]$\\rangle=0.29\\pm0.01$ for typical Milky Way GCs. We compare other abundance ratios with those of Local Group GCs and find indications for systematic differences in N and Ca abundance. The specific frequencies, $S_N$, of our sample galaxies are in line with the predictions of a simple mass-loss model for dwarf galaxies and compare well with $S_N$ values found for dwarf galaxies in nearby galaxy clusters. ", "introduction": "\\label{intro} The hierarchical structure formation scenario predicts that dwarf galaxies are the first systems to form in the Universe \\citep{peebles68}, and that more massive galaxies form through dissipative processes from these smaller sub-units. The involved physical mechanisms of this sequence depend on the density and mass of the parent dark matter halo, in the sense that more massive halos initiate star formation at earlier epochs and form their stars at a faster rate \\citep[e.g.][]{peebles02, renzini06, ellis07}. Because of this environmental gradient, we expect that dwarf galaxies in the field formed the first stellar population relatively late and at a lower pace compared to their counterparts in dense galaxy clusters. In other words, the difference in age and chemical composition between the oldest stellar populations in cluster and field dwarf galaxies should reflect the delay in the onset of structure formation in these two environments. The task of measuring the age and chemical composition of the oldest stellar populations in distant dwarf galaxies from their integrated light is very challenging. An alternative approach is to investigate the oldest globular clusters (GCs) that are found in dwarf galaxies. Several photometric surveys of extragalactic GCs in dwarf galaxies outside the Local Group have been performed in the past decade \\citep[see review by][]{miller06}. However, only a handful of those were followed up with 8--10m-class telescopes to derive spectroscopic ages and chemical composition. Observations of galaxies in groups and clusters provide more and more evidence that environment is a major factor influencing the process of GC formation \\citep[e.g.][]{west93, tully02, grebel03, miller98}.~Recent progress in modeling the assembly history of GC systems in massive elliptical galaxies suggests that a significant fraction of metal-poor GCs were accreted from dwarf satellites at later times compared to the number of GCs initially formed in the parent galaxy halo \\citep{ppm07}. These results underline the ideas put forward in the work of \\cite{forte82} and \\cite{muzzio87} as well as the models of \\cite{cote98, cote02, hilker99}, who suggested that the rich GC systems of massive galaxies may be the result of significant GC accretion through tidal stripping of less massive systems. Spectroscopic studies of a few GCs in cluster and field dwarf galaxies showed that most of these systems host at least some old GCs with ages $t\\ga10$ Gyr \\citep{puzia00, sharina03, strader03a, strader03b, beasley06, conselice06}.~Although today's accuracy of relative spectroscopic age determinations ($\\Delta t/t \\approx 0.2-0.3$) is not sufficient to resolve the expected delay of $\\sim\\!0.5-4$ Gyr in the onset of star-formation between cluster and field environment \\citep[depending on cosmology, ionizing source population, ionization feedback efficiency, etc., see][]{kauffmann96, treu05, thomas05, delucia06, clemens06}, the old ages combined with information on abundance ratios can provide a powerful tool to decide whether stellar populations in field dwarf galaxies followed the same early enrichment history as their analogs in denser environments. Furthermore, any difference in GC chemical composition between dwarf and more massive galaxies opens an attractive way of chemically tagging accreted sub-populations in massive halos and, therefore, enables us to quantify the mass accretion history of galaxies, a task that for old galaxies is infeasible from studies of the diffuse galaxy light alone. Similar ideas have been formulated for stellar populations that make up the diffuse component of the most nearby galaxies, which are close enough for high-resolution spectroscopy of individual stars \\citep{fh02, geisler07}. The obvious advantage of GC systems is that their spectra can be observed out to about 10 times greater distances. In this work we analyze spectroscopic observations of GCs in dwarf irregular (dIrr) and dwarf spheroidal galaxies (dSph) in the field/group environment. Our sample consists of systems that are representatives of the lowest-mass bin of the Local Volume (LV) galaxy population, limited to distances $D<10$ Mpc. We refer to \\cite{karachentseva85} and \\cite{grebel99} for a morphological type definition of our sample galaxies. The paper is organised as follows. In Section~\\ref{observations} we describe our observations and data reduction as well as the methods of measuring radial velocities. Section~\\ref{analysis} summarizes the measurement of Lick line indices, their calibration, and the determination of spectroscopic ages, metallicities, and abundance ratios. Section~\\ref{discussion} is devoted to the discussion of our results. Conclusions are presented in section~\\ref{conclusion}. \\begin{deluxetable*}{lccccccclcr}[!ht] \\tabletypesize{\\scriptsize} \\tablecaption{Properties of sample dwarf galaxies \\label{dwgprop}} \\tablewidth{0pt} \\tablehead{ \\colhead{Galaxy} & \\colhead{RA (J2000)}& \\colhead{DEC (J2000)} & \\colhead{$ \\mu_0$} & \\colhead{$D_{\\rm MD}$} & \\colhead{$A_{B}$} & \\colhead{$B$} & \\colhead{$B-I$} & \\colhead{$M_V$} & \\colhead{$N_{\\rm GC}$} & \\colhead{$S_{N}$} } \\startdata KK211, AM1339-445 & 13 42 06 & $-$45 13 18& 27.77 & 0.25 & 0.477 & 16.3$\\pm$0.2 & 1.8$\\pm$0.2 & $-$12.58& 2 & 18.6\\\\ KK221 & 13 48 46 & $-$46 59 49& 28.00 & 0.50 & 0.596 & 17.3$\\pm$0.4 & 2.0$\\pm$0.4 & $-$11.96& 6 & 95.1\\\\ KK084, UA200, KDG65 & 10 05 34 & $-$07 44 57& 29.93 & 0.03 & 0.205 & 16.4$\\pm$0.2 & 1.7$\\pm$0.2 & $-$14.40& 7 & 10.4\\\\ UGC3755, PGC020445 & 07 13 52 & $+$10 31 19& 29.35 &$\\sim5$ & 0.384 & 14.1$\\pm$0.2 & 1.1$\\pm$0.2 & $-$16.12& 32 & 11.4\\\\\\smallskip E490-017, PGC019337 & 06 37 57 & $-$25 59 59& 28.13 & 4.50 & 0.377 & 14.1$\\pm$0.2 & 1.1$\\pm$0.2 & $-$14.90& 5 & 5.4 \\\\ \\enddata \\tablecomments{Columns contain the following data: (1) galaxy name, (2) equatorial coordinates, (3) and (4) are the distance modulus and distance from the nearest bright galaxy in Mpc from \\cite{kara04}, and from \\cite{tully05} for UGC3755, (5) reddening from \\cite{sch98} in the B-band, (6) is the integrated $B$ magnitude, and (7) the integrated $B-I$ color, derived from surface photometry on the VLT-FORS2 images in this work (see Appendix~\\ref{sbprof}), (8) absolute $V$ magnitude from SPM05, (9) number of GCs according to SPM05 and this paper, (10) specific frequency, $S_N=N_{\\rm GC}10^{0.4(M_V+15)}$ \\citep{harris81}.} \\end{deluxetable*} ", "conclusions": "\\label{conclusion} Numerous photometric and spectroscopic studies of globular clusters in Virgo and Fornax cluster dwarf galaxies have been undertaken in the last years, which targeted bright dwarf galaxies down to $M_V\\approx-15$ mag \\citep[see][]{miller06}. Due to observational selection effects dwarf galaxies fainter than this are missed at distances of $D \\approx 17$ Mpc. Faint LSB dwarf galaxies down to $M_V \\approx -12$ mag have long been thought to be free of globular clusters, because they have insufficient mass. Our HST/WFPC2 survey of low-mass dwarf galaxies (SPM05), situated at distances $2-6$ Mpc in the Local Volume, revealed a rich population of globular cluster candidates (GCCs). In this work, we observed five of these galaxies with the VLT/FORS2 spectrograph in MXU mode and found that all targeted GCCs except one are genuine globular clusters. We could also confirm five additional globular clusters in our sample galaxies. Two clusters appear to be the nuclei of KK84 and KK211. The confirmed globular clusters are in general old and metal-poor, and show a range of \\afe\\ ratios. The mean $\\langle$[$\\alpha$/Fe]$\\rangle=0.19\\pm0.04$ that was determined with the $\\chi^{2}$ minimization technique and $0.18\\pm0.12$ dex which was computed using the iterative approach appears slightly lower than the mean $\\langle$[$\\alpha$/Fe]$\\rangle=0.29\\pm0.01$ for typical Milky Way clusters. Globular clusters in the two isolated, relatively bright dwarf galaxies UGC~3755 and ESO~490-17 show a wide range of ages from 1 to 9 Gyr, and imply extended star formation histories in these galaxies. This goes in hand with the measured low \\afe\\ ratios for the younger clusters and is consistent with low intensity star bursts. The oldest clusters with the highest \\afe\\ are found in KK84, a companion of NGC~3115. Other chemical abundances indicate potentially interesting differences between globular clusters in dwarf and more massive galaxies and, if confirmed, would facilitate the quantification of the accreted mass in rich GC systems of massive early-type galaxies." }, "0710/0710.4143_arXiv.txt": { "abstract": "The gravitational magnification and demagnification of Type Ia supernovae (SNe) modify their positions on the Hubble diagram, shifting the distance estimates from the underlying luminosity-distance relation. This can introduce a systematic uncertainty in the dark energy equation of state (EOS) estimated from SNe, although this systematic is expected to average away for sufficiently large data sets. Using mock SN samples over the redshift range $0 < z \\leq 1.7$ we quantify the lensing bias. We find that the bias on the dark energy EOS is less than half a percent for large datasets ($\\gtrsim$ 2,000 SNe). However, if highly magnified events (SNe deviating by more than 2.5$\\sigma$) are systematically removed from the analysis, the bias increases to $\\sim$ 0.8\\%. Given that the EOS parameters measured from such a sample have a 1$\\sigma$ uncertainty of 10\\%, the systematic bias related to lensing in SN data out to $z \\sim 1.7$ can be safely ignored in future cosmological measurements. ", "introduction": "\\label{sec:introduction} Since the discovery of the accelerating expansion of the universe~\\citep{Riess:98, Perl:99, Knop:03, Riess:04}, the quest to understand the physics responsible for this acceleration has been one of the major challenges of cosmology. At present the dominant explanation entails an additional energy density to the universe called dark energy. The physics of dark energy is generally described in terms of its equation of state (EOS), the ratio of its pressure to density. In some models this quantity can vary with redshift. While there exist a variety of probes to explore the nature of dark energy, one of the most compelling entails the use of type Ia supernovae to map the Hubble diagram, and thereby directly determine the expansion history of the universe. With increasing sample sizes, SN distances can potentially provide multiple independent estimates of the EOS when binned in redshift \\citep{Huterer:05,Sullivan:07a,Sullivan:07b}. Several present and future SN surveys, such as SNLS \\citep{snls} and the Joint Dark Energy Mission (JDEM), are aimed at constraining the value of the dark energy EOS to better than 10\\%. Although SNe have been shown to be good standardizable candles, the distance estimate to a given SN is degraded due to gravitational lensing of its flux \\citep{Frieman:97, Wambsganss:97, HolzWald:98}. The lensing becomes more prominent as we observe SNe out to higher redshift, with the extra dispersion induced by lensing becoming comparable to the intrinsic dispersion (of $\\sim0.1$ magnitudes) at $z \\gtrsim 1.2$ \\citep{Holz:05}. In addition to this dispersion, which leads to an increase in the error associated with distance estimate to each individual supernova, lensing also correlates distance estimates of nearby SNe on the sky, since the lines-of-sight pass through correlated foreground large-scale structure \\citep{Cooray:05,Hui:06}. Although this correlation error cannot be statistically eliminated by increasing the number of SNe in the Hubble diagram, the errors can be controlled by conducting sufficiently wide-area ($> 5\\mbox{ deg}^2$) searches for SNe (in lieu of small-area pencil-beam surveys). In addition to the statistical covariance of SN distance estimates, gravitational lensing also introduces systematic uncertainties in the Hubble diagram by introducing a non-Gaussian dispersion in the observed luminosities of distant SNe. Since lensing conserves the total number of photons, this systematic bias averages away if sufficiently large numbers of SNe per redshift bin are observed. In this case the average flux of the many magnified and demagnified SNe converges on the unlensed value \\citep{Holz:05}. Nonetheless, even with thousands of SNe in the total sample it is possible that the averaging remains insufficient, given that one may need to bin the Hubble diagram at very small redshift intervals to improve sensitivity to the EOS. Furthermore, SNe at higher redshifts are more likely to be significantly lensed. If ``obvious'' outliers to the Hubble diagram are removed from the sample, this introduces an important bias in cosmological parameter determination, and can lead to systematic errors in the determination of the dark energy EOS. In this paper we quantify the bias introduced in the estimation of the dark energy EOS due to weak lensing of supernova flux. We consider the effects due to the non-Gaussian nature of the lensing magnification distributions~\\citep{Wang:02}, performing Monte-Carlo simulations by creating mock datasets for future JDEM-like surveys. The paper is organized as follows: In $\\S$2.1 we discuss our parameterization of the dark energy EOS, $\\S$2.2 discusses gravitational lensing, and $\\S$3 is an in-depth description of our methodology. We present our results in $\\S$4. ", "conclusions": "\\label{sec:discussion} We first present our results for the magnitude case, as described in \\S\\ref{sec:mag}. The left panel of Figure~\\ref{fig:2} shows histograms of the best-fit values of $w_0$ from the likelihood analysis, after marginalizing over $w_a$ and $h$. The empty histogram, which peaks at -1.009 (marked with a vertical dot-dashed line), is for the model with 300 SNe. The shaded histogram, representing the 2,000 SN case, peaks at -1.003 (vertical dashed line), while the hatched histogram representing the 10,000 SN case has its peak at -1.002 (vertical solid line). These distributions have 1$\\sigma$ widths of 0.016, 0.006, and 0.003, respectively. This scatter is primarily due to the intrinsic uncertainty associated with absolute calibration, and is not dominated by lensing. Without the inclusion of lensing, however, the distributions peak at exactly -1, and show no bias. The shifted mode gives us a rough idea of the bias to be expected, on average, due to lensing. We find that 68\\% of the time a random sample of 300 SNe will have an estimated value for $w_0$ within 3\\% of its fiducial value, and this drops to 0.5\\% when a sample size of 10,000 SNe is considered. The right panel of Figure~\\ref{fig:2} shows the same distributions as the left panel, but this time using the flux-averaging technique instead of averaging over magnitudes. The empty, shaded, and hatched histograms peak at -1.007, -1.003, and -1.001, respectively, showing the mean bias for the 300, 2,000 and 10,000 SN cases (marked with dot-dashed, dashed, and solid vertical lines). With flux-averaging, we expect that 68\\% of the time a random sample of 300 SNe will yield a value of $w_0$ within 2.5\\% of the fiducial value, and within 0.5\\% for a sample of 10,000 SNe. The 1$\\sigma$ parameter uncertainty on $w_0$ ranges from the 20\\% level (for 300 SNe) to less than 5\\% (for 10,000 SNe), dwarfing the bias due to lensing. Thus, we need not be concerned about lensing degradation of dark energy parameter estimation for future {\\it JDEM}-like surveys. We note, however, that our estimated bias on the EOS is larger than the lensing bias of $w<0.001$ quoted in Table 7 of \\citet{Wood:07}. This is not surprising, given their use of the simple Gaussian approximation to lensing from \\citet{Holz:05}, which is less effective for low statistics. Nonetheless, we agree with their conclusion that lensing is negligible. A similar conclusion was also reached by \\citet{Martel:07} who used a compilation of 230 Type Ia SNe \\citep{Tonry:03} in the redshift range $0 < z < 1.8$ to show that the lensing errors are small compared to the intrinsic SNe errors. \\begin{figure}[!tb] \\begin{center} \\includegraphics[scale=0.5]{f3.eps} \\caption{Histograms showing the distribution of the values obtained for $w_0$ after marginalizing over $w_a$ and $h$ for the 2,000 SN case (using the flux-averaging technique). The shaded histogram assumes that the full sample of 2,000 SNe is used for parameter estimation. The hatched histogram shows the shift when outliers (SNe that are shifted above or below the Hubble diagram by more than 25\\% on either side) are removed from the sample. The bias in the distribution is due to the removal of highly-magnified lensing events from the sample.} \\label{fig:3} \\end{center} \\end{figure} We now discuss the bias which arises if anomalous SNe are removed from the sample. Gravitational lensing causes some SNe to be highly magnified, and it is conceivable that these ``obvious'' outliers are subsequently removed from the analysis. In this case the mean of the sample will be shifted away from the true underlying Hubble diagram, and a bias will be introduced in the best-fit parameters. To quantify this effect, we remove SNe which deviate from the expected mean luminosity-distance relation in the Hubble diagram by more than 25\\% (corresponding roughly to a $2.5\\sigma$ outlier). The SN scatter is a result of the convolution of the intrinsic error (Gaussian in flux of width 0.1) and the lensing PDF, and the outlier cutoff leads to a removal of $\\sim$ 50 SNe out of the 2,000 SNe. These outliers are preferentially magnified, due to the strong lensing tail of the magnification distributions. The demagnification tail is cut off by the empty-beam lensing limit, and therefore isn't as prominent. The hatched histogram in Figure~\\ref{fig:3} shows the distribution when events with convolved error greater than 2.5$\\sigma$ are removed. The vertical dot-dashed line at -1.0075 shows the average value of $w_0$ obtained in this case, representing a bias in the estimate of $w_0$ roughly three times larger than when the full 2,000 SNe are analyzed (shown by shaded histogram). This bias is a result of cutting off the high magnification tail of the distribution, and thus shifting the data towards a net dimming of observed SNe, leading to a more negative value of $w_0$. We also apply a cutoff at $3\\sigma$, in addition to the $2.5\\sigma$ discussed above. This results in a removal of $\\sim20$ SNe on average, for each 2,000 SN sample, and leads to a bias of $\\sim0.6$\\%. Any arbitrary cut on the (non-Gaussian) convolved (lensing + intrinsic) sample leads to a net bias in the distance relation, and even for large outliers and large SN samples, this can lead to percent-level bias in the best-fit values for $w_0$. To summarize, we have quantified the effect of weak gravitational lensing on the estimation of dark energy EOS from type Ia supernova observations. With generated mock samples of 2,000 SNe distributed uniformly in redshift up to z$\\sim$1.7 (as expected in future surveys like {\\it JDEM}), we have shown that the bias in parameter estimation due to lensing is less than 1$\\%$ (which is well within the 1$\\sigma$ uncertainty expected for these missions). Analyzing the data in flux or magnitude does not alter this result. If lensed supernovae that are highly magnified (such that the convolved error is more than 25\\% from the underlying Hubble diagram) are systematically removed from the sample, we find that the bias increases by a factor of almost three. Thus, so long as all observed SNe are used in the Hubble diagram, including ones that are highly magnified, the bias due to lensing in the estimate of the dark energy EOS will be significantly less than the 1$\\sigma$ uncertainty. Even for a post-{\\it JDEM} program with 10,000 SNe, lensing bias can be safely ignored." }, "0710/0710.1634_arXiv.txt": { "abstract": "We present 107 new epochs of optical monitoring data for the four brightest images of the gravitational lens SDSS J1004+4112 observed between October 2006 and June 2007. Combining this data with the previously obtained light curves, we determine the time delays between images A, B and C. We confirm our previous measurement finding that A leads B by $\\Delta t_{BA}=40.6\\pm1.8$~days, and find that image C leads image A by $\\Delta\\tau_{CA}=821.6\\pm2.1$ days. The lower limit on the remaining delay is that image D lags image A by $\\Delta\\tau_{AD}>1250$ days. Based on the microlensing of images A and B we estimate that the accretion disk size at a rest wavelength of 2300\\AA\\ is $10^{14.8\\pm0.3}$~cm for a disk inclination of $\\cos{i}=1/2$, which is consistent with the microlensing disk size-black hole mass correlation function given our estimate of the black hole mass from the MgII line width of $\\log M_{BH}/M_\\odot=8.44\\pm0.14$. The long delays allow us to fill in the seasonal gaps and assemble a continuous, densely sampled light curve spanning 5.7 years whose variability implies a structure function with a logarithmic slope of $\\gamma = 0.35\\pm0.02$. As C is the leading image, sharp features in the C light curve can be intensively studied 2.3 years later in the A/B pair, potentially allowing detailed reverberation mapping studies of a quasar at minimal cost. ", "introduction": "The quasar SDSS~J1004+4112 at $z_s=1.734$ is split into five images by an intervening galaxy cluster at $z_l=0.68$ \\cite{inada,inada2,oguri}. With a maximum image separation of $14\\farcs62$, it is a rare example of a quasar gravitationally lensed by a cluster \\cite{wambsganss,inada3}. One of the most interesting applications of this system is to use the time delays between the lensed images to study the structure of the cluster. If we assume the Hubble constant is known, then the delays break the primary model degeneracy of lensing studies (the ``mass sheet degeneracy''), and the delay ratios constrain the structure even if the Hubble constant is unknown. After its discovery, several groups modeled the expected time delays in SDSS J1004+4112 and their dependence on the mean mass profile of the cluster \\cite{kawano,oguri,williams}. When we measured the shortest delay in the system, between images A and B, we found a longer delay than predicted by the models (Fohlmeister et al. 2007, hereafter Paper I) where the discrepancy probably arose because the models included the cD galaxy and the cluster halo but neglected the significant perturbations from the member galaxies. As we measure the longer delays, where the cluster potential should be relatively more important than for the merging A/B image pair, we would not expect cluster substructures to play as important a role. We also expect this lens to have a fairly short time scale for microlensing variability created by stars either in the intracluster medium or in galaxies near the images. The internal velocities of a cluster are much higher than in a galaxy (700~km/s versus 200~km/s), and SDSS~J1004+4112's position on the sky is almost orthogonal to the CMB dipole (Kogut et al. 1993), giving the observer a projected motion on the lens plane of almost 300~km/s. In Paper I, we detected microlensing of the continuum emission of the A/B images in Paper I and there is also evidence for microlensing of the CIV broad line \\cite{richards,lamer,gomez}. Once we have measured the time delays we can remove the intrinsic quasar variability and use the microlensing variability to estimate the mean stellar mass and stellar surface density, the transverse velocities, and the structure of the quasar source \\cite{gilmerino,mortonson,poin,morgan}. Finally, we note that SDSS~J1004+4112 could be an ideal laboratory for studying correlations in the intrinsic variability of quasars. With, image C leading images A and B by 2.3 years, sharp variations in image C can be used to plan intensive monitoring of images A and B to measure the response times as a function of wavelength (e.g. Kaspi et al. 2007), with the additional advantage that the delay between A and B provides redundancies that protect against weather, the Moon and the Sun. The long delays between the images also mean that seasonal gaps are completely filled, and we can examine the structure function of the variability with a densely-sampled, gap-free light curve (modulo corrections for microlensing). Such data generally do not exist, since most time variability data for quasars (other than nearby reverberation mapping targets, e.g. Peterson et al. 2004) have very sparse sampling (e.g. Hawkins 2007 on long time scales for a small number of objects or Vanden Berk et al. 2004 on shorter time scales for many objects). In Paper I \\cite{fohli} we presented three years of optical monitoring data for the four brightest images of SDSS J1004+4112 spanning 1000 days from December 2003 to June 2006. The fifth quasar image, E, is too faint to be detected in our observations. We measured the time delay between the A and B image pair to be $\\Delta\\tau_{BA}=38.4\\pm2.0$ days. While larger separation lenses tend to have longer time delays, for these two images the propagation time difference is small, because they form a close image pair ($3\\farcs8$) from the source lying close to a fold caustic. For the more widely separated C and D images we could only estimate lower limits on the delays of 560 and 800 days relative to image B and A. In this paper we present the 107 new optical monitoring epochs for the 2006/2007 season in \\S2. When combined with our previous data we have light curves spanning 1250 days that allow us to measure the AC delay in \\S3. In \\S4 we use the microlensing variability of the A/B images to measure the size of the quasar accretion disk, and in \\S5 we measure the structure function of the intrinsic variability. We discuss the future prospects for exploiting this system in \\S6. \\begin{figure*}[t] \\centering \\includegraphics[bb= 30 100 520 700, width=10cm,angle=0,clip]{f1.ps} \\caption{ Light curves of the A, B, C and D images of the quasar SDSS J1004+4112 from December 2003 to June 2007. Images C and D have been offset by 0.3 and 1.0 mag, respectively, in order to avoid overlap. We present a running average of one point every 5 days averaged over $\\pm7$ days to emphasize trends and to avoid confusion by noise. \\label{lcurve}} \\end{figure*} ", "conclusions": "We present a fourth season of monitoring data for the four bright images of the five image gravitational lens system SDSS J1004+4112. We confirm our previous estimate for the time delay between the merging A/B pair, finding that B leads A by $40.6\\pm1.8$ days. We measure the delay for image C for the first time, finding that it leads image A by $821.6\\pm2.1$~days. We note that this is nearly twice the longest previously measured delay (the 417 day delay in Q0957+561 \\cite{schild, kundic}). We find a lower bound that D lags A by more than approximately 1250~days. Our current mass model predicts that D lags A by approximately 2000 days, which is consistent with the present limit. The fractional uncertainties in the AB delay are still dominated by sampling and microlensing, while the fractional uncertainties in the AC delay are dominated by cosmic variance due to density fluctuations along the line of sight rather than our measurement uncertainties of 0.3\\% (e.g. Barkana 1996). A detailed model of this system, including the constraints from the multiply imaged, higher redshift arcs (Sharon et al. 2005), the X-ray measurements \\cite{ota,lamer} and a detailed understanding of the uncertainties will be a challenge. We lack a completely satisfactory model for the system at present, in the sense that the modeling process is extraordinarily slow due to the ability of the gravitational potentials associated with the cluster member galaxies to generate additional but undetected images of the quasar, making it impossible to carry out a reliable model survey. The record of models for this system is discouraging. As we noted in Paper I, all three model studies (Oguri et al. 2004; Williams \\& Saha 2004; Kawano \\& Oguri 2006) generically predicted shorter AB delays than the observed 40 days, and that this could be plausibly explained by the absence of substructure (i.e. galaxies) in the potential models. The longer AB-C and AB-D delays should be less sensitive to substructure. Oguri et al. (2004) do not include an estimate of the AB-C delays and have A-D delays consistent with our present limits. The range of B-C delays in Williams \\& Saha (2004) is consistent with our measurement of 820 days, but they predict AD delays shorter than our current lower bound of 1250~days. Kawano \\& Oguri (2006) predict a range for the longer delays over a broad range of mass distributions, none of which match our delays in detail. However, models with sufficiently long C-B delays generally have C-D delays long enough to agree with our present limits. Based on our present mass model we used the microlensing between the A and B images to make an estimate of the size of the quasar accretion disk at 2300\\AA\\ in the quasar rest frame. If we convert this to the expected size at 2500\\AA\\ assuming the $R_\\lambda \\propto \\lambda^{4/3}$ scaling for a thin disk and assume the mean disk inclination $\\cos (i)=1/2$ the scale on which the disk temperature matches the photon energy is $R_{2500\\AA}=10^{15.0\\pm0.3}$~cm. Comparisons to other disk models should use the half-light radius which is $2.44$ times larger. Based on the quasar MgII emission line width we estimate that the black hole mass is $10^{8.4\\pm0.2} M_\\odot$. For this mass, the microlensing accretion disk size-black hole mass correlation found by Morgan et al. (2007) predicts that $R_{2500\\AA}=10^{15.3}$~cm, which is in broad agreement with the measurement. Further observations, the inclusion of additional images, and monitoring in multiple bands should improve these measurements and potentially allow us to determine the mean surface density in stars near the images $\\kappa_*$ and their average mass $\\langle M\\rangle$. Similarly, the ability to construct continuous light curves of the intrinsic variability and to use image C to provide early warning of sharp flux changes that can then be intensively monitored in images A and B may make this system a good candidate for applying reverberation mapping techniques to a massive, luminous quasar. At present, we already see that the system has a structure function typical of quasars." }, "0710/0710.3631_arXiv.txt": { "abstract": "We use high-resolution, three-dimensional hydrodynamic simulations to study the hydrodynamic and gravitational interaction between stellar companions embedded within a differentially rotating common envelope. We evaluate the contributions of the nonaxisymmetric gravitational tides and ram pressure forces to the drag force and, hence, to the dissipation rate and the mass accumulated onto the stellar companion. We find that the gravitational drag dominates the hydrodynamic drag during the inspiral phase, implying that a simple prescription based on a gravitational capture radius significantly underestimates the dissipation rate and overestimates the inspiral decay timescale. Although the mass accretion rate fluctuates significantly, we observe a secular trend leading to an effective rate that is significantly less than the rate based on a gravitational capture radius. We discuss the implications of these results within the context of accretion by compact objects in the common-envelope phase. ", "introduction": "To understand the evolution of binary star systems, it is essential to analyze the interactions between their stellar components. Examples of such influences include the spin-orbit tidal interaction and mass transfer, as well as interactions that result in the loss of mass and angular momentum. Equally important are the interactions of stars orbiting about their common center of mass within a differentially rotating common envelope. It is generally accepted that such an evolutionary stage is essential for the formation of short-period binary systems containing compact objects (see, e.g., Iben \\& Livio 1993 and Taam \\& Sandquist 2000). In this case, the interaction determines the orbital evolution of the system and the conditions under which the common envelope is ejected, leading to the survival of a remnant binary system or to a merger that forms a rapidly rotating single star. The amount of mass and angular momentum accreted by the inspiralling components during this phase also has direct implications for the properties of the compact object population in binary systems. Lacking multidimensional hydrodynamical simulations of the common-envelope phase, the initial numerical and semi-analytical studies of the problem used simple prescriptions for the stellar interactions based on the pioneering work by Hoyle \\& Lyttleton (1939) and Bondi \\& Hoyle (1944), as generalized by Bondi (1952). These seminal studies focused on the idealized problem of the capture of matter by a gravitating point object moving supersonically with respect to a uniform medium. In this framework, a gravitational capture radius, $\\rcap$, plays an important role in determining the rates of mass accretion and energy dissipation. $\\rcap$ is given by \\begin{equation} \\label{Eqn:rcap} \\rcap = {2 G M \\over \\vrel^2 + \\cs^2}\\ , \\end{equation} where $M$ is the mass of the gravitating object, $\\vrel$ is the velocity of the object with respect to the medium, and $\\cs$ is the local speed of sound. When a density gradient with scale height $H$ is present, the effective accretion radius $\\racc$ is (Dodd \\& McCrea 1952) \\begin{equation} \\label{Eqn:racc} \\racc = {\\rcap \\over 1 + (\\rcap/2H)^2}\\ . \\end{equation} The energy dissipation rate is then $L_{\\rm d} \\approx \\pi \\racc^2 \\rho \\vrel^3$, where $\\rho$ is the upstream density. To improve on these estimates, hydrodynamic effects were approximated analytically by Ruderman \\& Spiegel (1971), Wolfson (1977), and Bisnovatyi-Kogan et al.\\ (1979) as well as numerically by Hunt (1971, 1979), Shara \\& Shaviv (1980), and Shima et al.\\ (1985). These early multidimensional simulations considered axisymmetric flow, and their results have been used to calibrate the energy loss rate. In particular, the drag coefficients obtained from such simulations (see, e.g., Shima et al.\\ 1985) have been used to estimate the rate of energy dissipation in the common envelope. However, many of the simplifying assumptions underlying these studies are inadequate for direct application to common-envelope interactions. The flow is nonaxisymmetric and distinctly nonuniform, reflecting the existence of velocity or density gradients (the density may span several scale heights within $\\racc$). The effect of relaxing these assumptions has been studied in two dimensions by Fryxell \\& Taam (1988) and Taam \\& Fryxell (1989) and in three dimensions by Sawada et al.\\ (1989) and Ruffert (1999). These studies could not encompass the full complexity of common-envelope interactions, since the envelope's self-gravity was ignored. Furthermore, because the companions move in elliptical orbits, their cores interact with matter that has already been affected in previous orbital phases. Thus, the state of the gas and its environment in these calculations must be regarded as highly idealized. Within the past decade, three-dimensional numerical studies of the common-envelope phase that have relaxed earlier geometrical assumptions have been carried out by Sandquist et al.\\ (1998, 2000), DeMarco et al.\\ (2003a,b), and Taam \\& Ricker (2006). Recently, we have carried out high-resolution adaptive mesh refinement (AMR) simulations of common-envelope evolution with effective resolutions of $2048^3$ (R. E. Taam \\& P. M. Ricker, in preparation), allowing the interaction of the stars within the common envelope to be examined and quantified. In this Letter we report on some results of our numerical studies. We focus on analyzing a single high-resolution simulation to determine the hydrodynamic and gravitational contributions to the drag forces affecting the orbital motion of the stellar components during the early inspiral phase. We also analyze the accumulation of mass by the stellar components within the common envelope to compare their magnitudes to estimates based on an accretion radius formalism. In \\S~2, we briefly describe the numerical method and our assumed model for the binary system. Descriptions of the method of analysis and the numerical results are presented in \\S~3. Finally, we summarize our results and comment on their possible implications for applications involving compact objects in short-period binary systems. ", "conclusions": "We have quantitatively described the interaction of stars within a common envelope on the basis of the analysis of the early inspiral stage of a $1.05 \\msun$ red giant and a $0.6 \\msun$ binary companion. The orbital decay is dominated by the nonaxisymmetric gravitational drag associated with the self-gravitating matter in the common envelope. On the basis of high-resolution three-dimensional hydrodynamic simulations, this drag is 1-2 orders of magnitude greater than the hydrodynamic drag. As a consequence, the orbital decay timescale is much shorter than that derived from analyses based on the Hoyle-Lyttleton-Bondi picture of accretion from a uniform medium by a gravitating point mass moving at supersonic speeds. In this latter description the gravitational drag associated with an accretion wake is not significantly larger than the hydrodynamic drag. The effect of long-range gravitational interactions is critical for reliable estimates of the orbital decay timescale and energy dissipation rate. In this picture the drag and the mass accretion rate are not as directly related as they are in the Hoyle-Lyttleton-Bondi-type description, because the dominant drag term is gravitational, rather than hydrodynamic, in origin. This decoupling is reflected in the much larger ratio $L_{\\rm d} / \\dot{M} \\sim 10^{15}-10^{16}$~cm$^2$~s$^{-2}$ measured in the simulation than expected from the gravitational capture radius formalism ($\\sim 10^{14}$~cm$^2$~s$^{-2}$). Although the mass accretion rates estimated from our simulation should be regarded as only indicative due to the lack of a detailed inner boundary treatment for the companion, the structure of the flow (dominated as it is by tidal effects) strongly suggests that the true mass accretion rate should be much smaller than the rate expected in the gravitational capture radius formalism. The effective capture radius, on the basis of the observed ambient density and relative velocity, is almost an order of magnitude smaller than the expected value. In any case, the expected value would lead to an unrealistic level of accretion over the common-envelope period. Furthermore, the inspiral time suggested by the early evolution of our three-dimensional simulation is much shorter than that found in earlier one- and two-dimensional calculations that are based on the Bondi-Hoyle-Lyttleton picture. Assuming that the discrepancy in mass accretion rate continues into the deep inspiral phase, the total accumulation of matter onto the companion should be much smaller than previously expected. These results would have little effect on the mass of an embedded main-sequence star because of its tendency to expand as a result of the high entropy within the common envelope (see Hjellming \\& Taam 1991); however, it can significantly affect the outcome for neutron stars within a common envelope. In particular, the estimated accretion rates exceed $10^{-3} \\mpy$, for which steady state accretion flows with neutrino losses are possible (Chevalier 1989; Houck \\& Chevalier 1991). At these hypercritical mass accretion rates, photons are trapped in the flow and the Eddington limit is not applicable. On the basis of this hypercritical accretion flow regime, Chevalier (1993) and Brown (1995) suggest that neutron stars embedded in the common envelope would accrete sufficient mass to form low-mass black holes (although see Chevalier 1996), and, hence, the formation of binary radio pulsars would require an evolutionary scenario involving progenitor stars of nearly equal mass (Brown 1995). However, the population synthesis of binary black holes and neutron stars by Belczynski et al.\\ (2002), including hypercritical accretion, resulted in an average accretion of $0.4 \\msun$. Such a high rate of mass accretion is inconsistent with the observed masses of binary radio pulsars ($\\sim 1.35 \\msun$; Thorsett \\& Chakrabarty 1999) and indicates the need for a reduction in accreted matter during the common-envelope phase (see also Belczynski et al.\\ 2007). Our calculations show that the necessary reduction may arise naturally as a result of a more realistic treatment of the common-envelope phase. Consequently, this reduction could also lead to a reduction in the number of low-mass black holes, depending on the maximum mass of neutron stars, resulting from the accretion induced collapse of massive accreting neutron stars in the common-envelope phase. Similarly, the mass accretion, which was found to be as large as several solar masses for black hole accretors, would also be reduced, thereby affecting the masses and spins of double black holes emerging from the common-envelope phase. Further investigations are planned to examine the generality of these results regarding mass accretion and to quantify the importance of these processes for determining the properties (mass and spin) and ultimate fate of the compact components in short-period binary system populations. Such studies are not only important for determining the masses of binary neutron star and black hole systems resulting from the common-envelope phase (Belczynski et al.\\ 2007), but also their orbital periods, which directly influence the expected merger rates of such binary populations as sources for gravitational wave detection in the advanced LIGO experiment." }, "0710/0710.3588_arXiv.txt": { "abstract": "We assess the relative merits of weak lensing surveys, using overlapping imaging data from the ground-based Subaru telescope and the Hubble Space Telescope (HST). Our tests complement similar studies undertaken with simulated data. From observations of 230,000 matched objects in the 2 square degree COSMOS field, we identify the limit at which faint galaxy shapes can be reliably measured from the ground. Our ground-based shear catalog achieves sub-percent calibration bias compared to high resolution space-based data, for galaxies brighter than $i^{\\prime}\\simeq$24.5 and with half-light radii larger than $1.8\\arcsec$. This selection corresponds to a surface density of 15 galaxies arcmin$^{-2}$ compared to $\\sim 71$ arcmin$^{-2}$ from space. On the other hand the survey speed of current ground-based facilities is much faster than that of HST, although this gain is mitigated by the increased depth of space-based imaging desirable for tomographic (3D) analyses. As an independent experiment, we also reconstruct the projected mass distribution in the COSMOS field using both data sets, and compare the derived cluster catalogs with those from $X$-ray observations. The ground-based catalog achieves a reasonable degree of completeness, with minimal contamination and no detected bias, for massive clusters at redshifts $0.2$ 100~MeV). It will be able to locate sources to positional accuracies of 30 arc seconds to 5 arc minutes. The precision of this instrument could well be enough to detect a heavy neutrino signal in the form of a small bump at $E\\sim 1$ GeV in the gamma spectrum, if a heavy neutrino with mass $\\sim$100 or 200~GeV would exist. There are also some other possible consequences of heavy neutrinos that may be worth investigating. The DM simulations could be used to estimate the spatial correlations that the gamma rays would have and to calculate a power spectrum for the heavy neutrinos. This could be interesting at least for masses $M_N\\sim 100$~GeV and $M_N\\sim 200$~GeV. The annihilation of the heavy neutrinos could also help to explain the reionization of the universe. Another possible interesting application of heavy neutrinos would be the large look-back time they provide \\citep{2006PhLB..639...14S}, with a decoupling temperature of $\\gtrsim 10^{13}$ K \\citep{1989NuPhB.317..647E}." }, "0710/0710.4966_arXiv.txt": { "abstract": "{ The EGRET excess of diffuse Galactic gamma rays shows all the features expected from dark matter annihilation (DMA): a spectral shape given by the fragmentation of mono-energetic quarks, which is the same in all sky directions and an intensity distribution of the excess expected from a standard dark matter halo, predicted by the rotation curve. From the EGRET excess one can predict the flux of antiprotons from DMA. However, how many antiprotons arrive at the detector strongly depends on the pro\\-pagation model. The conventional isotropic propagation models trap the antiprotons in the Galaxy leading to a local antiproton flux far above the observed flux. According to Bergstr\\\"{o}m et. al. this excludes the DMA interpretation of the EGRET excess. Here it is shown that more realistic anisotropic propagation models, in which most antiprotons escape by fast transport in the z-direction, are consistent with the B/C ratio, the antiproton flux and the EGRET excess from DMA. } % ", "introduction": "\\label{intro} The interpretation of the observed EGRET excess of diffuse Galactic gamma rays as Dark Matter annihilation (DMA) (see \\cite{us} or contributions by W. de Boer, C. Sander and M. Weber, this volume) could be a first hint at the nature of dark matter. The excess was observed in all sky directions. From the spectral shape of the excess the WIMP mass was constrained to be between 50 and 100 GeV and from the distribution of the excess in the sky the Dark Matter (DM) halo profile was obtained. One of the most important criticisms of this analysis was a paper by Bergstr\\\"{o}m et. al. \\cite{bergstrom1} claiming that the antiproton flux from DMA would be an order of magnitude higher than the observed antiproton flux. They used a conventional propagation model assuming the propagation of charged particles to be the same in the halo and the disk. However, the propagation in the halo (perpendicular to the disk) can be much faster than the propagation in the disk \\cite{breit}. In this paper we show that the local antiproton flux from DMA can be strongly reduced in an anisotropic propagation model and that the DMA interpretation of the EGRET excess can by no means be excluded by Galactic antiprotons. In section \\ref{CM} we discuss the pro\\-blems of the isotropic model for cosmic ray transport leading to the fact that our galaxy can work as a large storage box for antiprotons. An anisotropic pro\\-pagation model, which simultaneously describes the \\\\ EGRET excess and and the observed local fluxes of charged cosmic rays is introduced in section \\ref{CP}. Section \\ref{conclusion} summarizes the results. ", "conclusions": "Tracing of charged particles in realistic models of the regular Galactic magnetic fields with a turbulent (small-scale) component has shown that CRs remember the regular field lines, even if the irregular component is of the same order of magnitude as the regular, thus leading to enhanced diffusion in $\\phi$ and $z$ (see Fig. A1 in \\cite{blasi}). With such an anisotropic propagation model the amount of antiprotons expected from DMA can be reduced by one to two orders of magnitude. Therefore the claim by \\cite{bergstrom1} that the DMA interpretation of the EGRET excess of diffuse Galactic gamma rays is excluded is strongly pro\\-pagation model dependent. It only applies to a pro\\-pagation model with isotropic diffusion. An anisotropic pro\\-pagation model with different pro\\-pagation in the halo and the disk can reconcile the EGRET excess with the antiproton flux and the ratios of secondary/primary and unstable/stable nuclei. Clearly the DMA search for light DM particles is propagation model dependend. Taking these uncertainties into account shows that DMA is a viable explanation of the EGRET excess of diffuse Galactic gamma rays, as shown in \\cite{us} and and can by no means be excluded by the antiproton flux predicted by a specific model. \\vspace*{5mm}" }, "0710/0710.0774_arXiv.txt": { "abstract": "In the third part of our photometric study of the star-forming region NGC~346/N~66 and its surrounding field in the Small Magellanic Cloud (SMC), we focus on the large number of low-mass pre-main sequence (PMS) stars revealed by the Hubble Space Telescope Observations with the Advanced Camera for Surveys. We investigate the origin of the observed broadening of the pre-main sequence population in the $V-I$, $V$ CMD. The most likely explanations are either the presence of differential reddening or an age spread among the young stars. Assuming the latter, simulations indicate that we cannot exclude the possibility that stars in NGC 346 might have formed in two distinct events occurring about 10 and 5 Myr ago, respectively. We find that the PMS stars are not homogeneously distributed across NGC 346, but instead are grouped in at least five different clusters. On spatial scales from 0.8$''$ to 8$''$ (0.24 to 2.4\\,pc at the distance of the SMC) the clustering of the PMS stars as computed by a two-point angular correlation function is self-similar with a power law slope $\\gamma \\approx -0.3$. The clustering properties are quite similar to Milky Way star forming regions like Orion OB or $\\rho$\\,Oph. Thus molecular cloud fragmentation in the SMC seems to proceed on the same spatial scales as in the Milky Way. This is remarkable given the differences in metallicity and hence dust content between SMC and Milky Way star forming regions. ", "introduction": "It is well known that Galactic OB associations also host large numbers of fainter, low-mass pre-main sequence (PMS) stars (e.g.\\ Preibisch \\& Zinnecker 1999; Sherry et al.\\ 2004; Brice\\~{n}o et al.\\ 2005). Low-mass PMS stars in stellar associations provide a longer-lasting record of the most recent star formation events than the short-lived high-mass stars. Large-scale surveys have identified hundreds of PMS stars in nearby OB associations (Brice\\~{n}o et al. 2007). In order to understand the star formation triggering and history, or the Initial Mass Function (IMF), one has to study both high- and low-mass stars in star forming regions. In many cases the low-mass populations of galactic OB associations cannot be easily distinguished from foreground main-sequence stars (e.g. Brice\\~{n}o et al. 2001). Recently PMS stars have been discovered for the first time in an extragalactic stellar association, which suffers much less from contamination by the galactic disk. {\\em Hubble Space Telescope} (HST) observations of the association LH\\,52 in the Large Magellanic Cloud (LMC) revealed $\\approx$500 PMS stars with masses down to 0.3\\,M$_\\odot$ (Gouliermis et al. 2006a). The investigation of individual extragalactic pre-main sequence stars opens a new field of study. As both high angular resolution and wide field-of-view are required, HST is the ideal observatory to carry out these studies. While HST observations of extragalactic associations cannot reach the detection limits achieved for local associations, they provide a unique opportunity for studying low-mass star formation in other galaxies. The OB association NGC~346 in the Small Magellanic Cloud (SMC) is located in the central part of the brightest {\\sc H ii} region in the SMC, named LHA~115-N~66 or in short N~66 (Henize 1956). With 33 spectroscopically confirmed O and B stars, NGC~346 hosts the largest sample of young, massive stars in the SMC (Walborn 1978; Walborn \\& Blades 1986; Niemela et al. 1986; Massey et al. 1989; Walborn et al. 2000; Evans et al. 2006). Photoionization models by Rela\\~{n}o et al. (2002) imply that these stars are a major source of ionizing flux for the surrounding diffuse interstellar medium (ISM). Recent imaging from the Wide-Field Channel (WFC) of the {\\em Advanced Camera for Surveys} (ACS) on-board HST (GO Program 10248; PI: A. Nota) revealed the PMS stellar content of the general region of NGC~346/N~66 down to the sub-solar mass regime (Nota et al.\\ 2006; Gouliermis et al.\\ 2006b -- hereafter Paper I -- ; Sabbi et al.\\ 2007). Nota et al.\\ (2006) suggest that all PMS stars in the association are the product of a single star formation event, taking place 3 to 5 Myr ago. The PMS distribution in the $V-I$, $V$ color-magnitude diagram (CMD), however, shows a prominent widening, which may be explained by an age spread of $\\approx$10 Myr. This raises the question {\\em if the PMS stars in NGC\\,346 are indeed the result of a single star formation event.} Sabbi et al.\\ suggest that the PMS population is mainly concentrated in a number of subclusters (three of them at the central part, where the association is located), which formed at the same time from the turbulence-driven density variations, and not following a sequential process. Star counts, however, reveal only a few compact PMS clusters with a significant number of stars in the area around the association. The main body of the association cannot be divided into separate subclusters (Paper I). {\\em Could the remote clusters be the product of a star formation event occurring before or after the event which triggered the formation of the association?} In the first part of our study of this extraordinary star-forming region (Paper I), we compiled our ACS photometry of almost 100,000 stars, and presented preliminary results on stellar types and their distinct spatial distributions. We confirmed the co-existence of massive OB stars and low-mass PMS stars in NGC 346. In the present paper, we explore in greater detail the properties of the low-mass PMS stars with a focus on clustered star formation in NGC~346/N~66. The paper is organized as follows: In \\S~2 we describe the datasets from HST/ACS, present the color-magnitude diagram, carry out a comparison of our photometry of the brightest stars with previous photometric studies, and discuss the reddening in the region. The broadening of the PMS distribution in the CMD, a thorough discussion of possible causes and explanations of this phenomenon as well as the spatial variations in the CMD are presented in \\S~3. In \\S~4 we present the clustering properties of the PMS stars and discuss them in terms of hierarchical star formation. Our analysis of H\\alp\\ observations with HST/ACS as well as previous {\\em Spitzer} Space Telescope results on Young Stellar Objects (YSOs) are presented in \\S~5 and \\S~6, respectively. In \\S~7 we summarize the results. ", "conclusions": "In this paper we present the results from our photometric study on the recent star formation history of N~66 the brightest {\\sc H~ii} region in the SMC, related to the OB association NGC~346, as it is recorded in the observed PMS population of the region. Our deep photometry revealed an extremely large number of low-mass PMS stars in the association and the surrounding region, easily distinguishable in the $V-I$, $V$ CMD. We show that factors such as reddening, binarity and variability can cause a broadening of the positions of these stars in the CMD of the association, wide enough so that they can be misinterpreted as the result of multi-epoch star formation, and not as the product of a single star formation event. We found that a modest reddening like the one we found for the general area of NGC~346 ($E(B-V)\\simeq 0.08$ mag) can make the PMS stars appear younger than what they actually are, and therefore an age of 10 Myr better fits the observed sequence of PMS stars. However, our results do not exclude the possibility of multi-epoch star formation in the area of the association if the reddening is even lower, as suggested from the OB stars of this region ($E(B-V)\\simeq 0.04$ mag). In this case, two star formation events 10 and 5 Myr ago can explain the observed broadening of the PMS stars due to age spread and factors such as reddening and binarity. No specific dependence of the estimated ages of the PMS stars to their loci within the association, as a signature of sequential star formation, was found. It has been previously suggested that three different generations of stars occurred through sequential star formation in the region of NGC~346/N~66 (Rubio et al. 2000) within the last 3 Myr (based on the age of the youngest OB stars) and that the PMS stars of NGC~346 represent a star formation event that took place 3 - 5 Myr ago (Nota et al. 2006). However, the study by Massey et al. (1989) based on the presence of evolved 15~M{\\solar} stars suggests that there might have been earlier star formation events in the region making NGC~346 as old as $\\sim$~12 Myr. In addition, there are recent indications that NGC~346 might host classical Be-type stars (Evans et al. 2006), and if so, the age of the association should be at least 10~Myr, a threshold given by Fabregat \\& Torrej{\\'o}n (2000) for classical Be-type stars to form. These results fit very well to the hypothesis that star formation in NGC~346 did already occur about 10~Myr ago, as our observations and simulations of the PMS stars suggest. From star counts based on our ACS photometry we identify at least five PMS clusters across the region, covering a range of ages. On spatial scales from 0.8$''$ to 8$''$ (0.24 to 2.4\\,pc at the distance of the SMC) the clustering of the PMS stars as computed by a two-point angular correlation function is self-similar with a power law slope $\\gamma \\approx -0.3$. The clustering properties are quite similar to Milky Way star-forming regions like the Orion OB association or $\\rho$\\,Oph. Thus molecular cloud fragmentation in the SMC seems to proceed on the same spatial scales as in the Milky Way. This is remarkable given the differences in metallicity and hence dust content between SMC and Milky Way star-forming regions. The youngest PMS stars are located mostly to the north of the bar of N~66, where three PMS clusters are identified. This area is also characterized by a high concentration of candidate YSOs (Simon et al. 2007), H\\alp-excess stars (found with our photometry), and IR-emission peaks (Rubio et al. 2000). This indicates that star formation probably still takes place in an arc-like feature, as it is outlined by the spatial distribution of these sources. In an accompanying letter, we combine these results with previous multi-wavelength studies of the region to investigate the star formation history, which helped to shape NGC~346/N~66 (Gouliermis et al. 2007b)." }, "0710/0710.0890_arXiv.txt": { "abstract": "A next-generation lunar laser ranging apparatus using the 3.5~m telescope at the Apache Point Observatory in southern New Mexico has begun science operation. APOLLO (the Apache Point Observatory Lunar Laser-ranging Operation) has achieved \\emph{one-millimeter} range precision to the moon which should lead to approximately one-order-of-magnitude improvements in several tests of fundamental properties of gravity. We briefly motivate the scientific goals, and then give a detailed discussion of the APOLLO instrumentation. ", "introduction": "\\subsection{Scientific Motivation\\label{sub:Scientific-Motivation}} A variety of observations and theoretical explorations---including the apparent acceleration of the expansion of the universe \\citep{sn1,sn2}, the possible existence of extra dimensions \\citep{extra_dim}, and attempts to reconcile quantum mechanics and gravity---provide motivation for improved tests of the fundamental aspects of gravity. Lunar Laser Ranging (LLR) currently provides the best tests of a number of gravitational phenomena \\citep{jgw-96,jgw-latest} such as: \\begin{itemize} \\item the strong equivalence principle (SEP): $\\eta\\approx5\\times10^{-4}$ sensitivity \\item time-rate-of-change of the gravitational constant: $\\dot{G}/G<10^{-12}$ yr$^{-1}$ \\item geodetic precession: 0.6\\% precision confirmation \\item deviations from the $1/r^{2}$ force law: $\\sim10^{-10}$ times the strength of gravity at $10^{8}$ meter scales \\end{itemize} LLR also tests other gravitational and mechanical phenomena, including for example gravitomagnetism \\citep{gravmag}, preferred frame effects \\citep{alpha-1,alpha-2}, and Newton's third law \\citep{newtonsthird}. LLR may also provide a window into the possible existence of extra-dimensions via cosmological dilution of gravity \\citep{lue,dvalimoon}. Besides the SEP, LLR tests the weak equivalence principle (WEP) at the level of $\\Delta a/a<1.3\\times10^{-13}$, but the LLR constraint is not competitive with laboratory tests. In addition, LLR is used to define coordinate systems, probe the lunar interior, and study geodynamics \\citep{dickey}. These constraints on gravity are based on about 35 years of LLR data, although the precision is dominated by the last $\\sim$15 years of data at 1--3~cm precision. APOLLO aims to improve tests of fundamental gravity by approximately an order-of-magnitude by producing range points accurate at the one-millimeter level. \\subsection{A Brief History of LLR\\label{sub:A-Brief-History}} The first accurate laser ranges to the moon followed the landing of the first retroreflector array on the Apollo 11 mission by less than two weeks (August 1, 1969). These were performed on the 3.0~meter telescope at the Lick Observatory. One month later, a second station using the 2.7~meter telescope at the McDonald Observatory began ranging to the moon \\citep{bender}. The operation at the Lick Observatory was designed for demonstration of initial acquisition, so that the scientifically relevant observations over the next decade came from the McDonald station, which used a ruby laser with 4~ns pulse width, firing at a repetition rate of about 0.3 Hz and $\\sim3$ J/pulse. This station routinely achieved 20~cm range precision, with a photon return rate as high as 0.2 photons per pulse, or 0.06 photons per second. A typical ``normal point''---a representative measurement for a run typically lasting tens of minutes---was constructed from approximately 20 photon returns. In the mid 1980's, the McDonald operation was transferred to a dedicated 0.76~m telescope (also used for satellite laser ranging) with a 200~ps Nd:YAG laser operating at 10~Hz and 150 mJ/pulse. This station is referred to as the McDonald Laser Ranging System: MLRS \\citep{mlrs}. At about the same time, a new station began operating in France at the Observatoire de la C\\^ote d'Azur (OCA) \\citep{oca}. Using a 1.5~meter telescope, a 70~ps Nd:YAG laser firing at 10~Hz and 75 mJ/pulse, this became the premier lunar ranging station in the world. In recent years, the MLRS and OCA stations have been the only contributors to lunar range data with typical return rates of 0.002 and 0.01 photons per pulse, respectively. Typical normal points from the two stations consist of 15 and 40 photons, respectively. Other efforts in LLR are described in \\citet{ep-llr}, and more detailed histories may be found in the preceding reference as well as in \\citet{bender,dickey}. \\subsection{Millimeter Requirements\\label{sub:Millimeter-Requirements}} The dominant source of random uncertainty in modern laser ranging systems has little to do with the system components, but rather comes from the varying orientation of the lunar retroreflector arrays. Although the arrays are nominally pointed within a degree of the mean earth position, variations in the lunar orientation---called libration---produce misalignments as large as 10 degrees, and typically around 7 degrees. This means the ranges between the earth and the individual array elements typically have a root-mean-square (RMS) spread of 15--36~mm, corresponding to about 100--240~ps of round-trip travel time. This dominates over uncertainties associated with the laser pulse width, and with jitter in the detector and timing electronics. A typical normal point containing 16 photons will therefore be limited to 4--9~mm range precision by the array orientation alone, though range residuals reported by analysis at the Jet Propulsion Laboratory tend to be larger than this. Reaching the one-millimeter precision goal demands at a minimum the collection of enough photons to achieve the appropriate statistical reduction. Assuming an ability to identify the centroid of $N$ measurements---each with uncertainty $\\sigma$---to a level of $\\sigma_{\\mathrm{net}}=\\sigma/\\sqrt{N}$, the uncertainty stemming from the retroreflector array orientation typically demands 225--1300 photons in the normal point to reach the one millimeter mark. Worst-case orientations push the individual photon uncertainty to 50~mm, demanding 2500 photons. This is far outside of the capabilities of the aforementioned LLR stations. We point out that any constant range bias is accommodated in the analysis, so that only \\emph{variations} in the range are important to the experiment. While adequate photon number is sufficient to reduce statistical uncertainty to the one-millimeter level, other sources of error could potentially limit the ultimate scientific capacity of LLR. Most importantly, the gravitational physics is sensitive to the center-of-mass separations of Earth and Moon, while one measures the distance between a telescope and reflectors that are confined to the body surfaces. The earth's surface in particular has a rich dynamic---experiencing diurnal solid-earth tides of 350~mm peak-to-peak amplitude, plus crustal loading from oceans, atmosphere, and ground water that can be several millimeters in amplitude. Moreover, the earth atmosphere imposes a propagation delay on the laser pulse, amounting to $\\sim$1.5~m of zenith delay at high-altitude sites. Satellite laser ranging, very long baseline interferometry, and other geodetic efforts must collectively contend with these same issues, for which accurate models have been produced. A good summary of these models is published by the International Earth Rotation and Reference Systems Service \\citep[IERS:][]{iers}. As an example of the state of these models, the long-standing atmospheric model by \\citet{marini-murray} has recently been replaced by a more accurate model \\citep{atmo1,atmo2}. The model differences for a high-altitude site are no more than 2~mm for sky elevation angles greater than 40 degrees---providing an indicative scale for the model accuracy. The primary input for this model is the atmospheric pressure at the site, as this represents a vertical integration of atmospheric density, which in turn is proportional to the deviation of the refractive index, $n$, from unity. Thus the zenith path delay, being an integration of $n-1$ along the path, is proportional to surface pressure under conditions of hydrostatic equilibrium. A mapping function translates zenith delay to delays for other sky angles. Measuring pressure to a part in 2000 (0.5~mbar) should therefore be sufficient to characterize the 1.5~m zenith delay at the one-millimeter level. Our experiment records atmospheric pressure to an accuracy of 0.1~mbar. The principal science signals from LLR appear at well-defined frequencies. For example, the equivalence principle signal is at the synodic period of 29.53 days, and even secular effects ($\\dot{G}$, precession) are seen via the comparative phases between periodic (monthly) components in the lunar orbit. Because many of the effects discussed in the preceding paragraphs are aperiodic, they will not mimic new physics. To the extent that these effects are not adequately modeled, they contribute either broadband noise or discrete ``signals\" at separable frequencies. The science output from APOLLO may be initially limited by model deficiencies. But APOLLO's substantial improvement in LLR precision, together with a high data rate that facilitates deliberate tests of the models, is likely to expose the nature of these deficiencies and therefore propel model development---as has been historically true for the LLR enterprise. Ultimately, we plan to supplement our LLR measurement with site displacement measurements from a superconducting gravimeter (not yet installed), in conjunction with a precision global positioning system installation as part of the EarthScope Plate Boundary Observatory (installed February 2007 as station P027). \\subsection{The APOLLO Contribution} APOLLO---operating at the Apache Point Observatory (APO)---provides a major improvement in lunar ranging capability. The combination of a 3.5~meter aperture and 1.1~arcsecond median image quality near zenith translates to a high photon return rate. Using a 90~ps FWHM (full-width at half-maximum) Nd:YAG laser operating at 20~Hz and 115~mJ/pulse, APOLLO obtains photon return rates approaching one photon per pulse, so that the requisite number of photons for one-millimeter normal points may be collected on few-minute timescales. To date, the best performance has been approximately 2500 return photons from the Apollo 15 array in a period of 8 minutes. The average photon return rate for this period is about 0.25 photons per shot, with peak rates of 0.6 photons per pulse. Approximately half of these photons arrived in multi-photon bundles, the largest containing eight photons. APOLLO brings LLR solidly into the multi-photon regime for the first time. This paper describes the physical implementation of the APOLLO apparatus, including descriptions of the optical and mechanical design, the electronics implementation, and system-level design. For early reports on APOLLO, see \\citet{iwlr12,spacepart,iwlr13,iwlr14}. For an analysis of our expected photon return rate, see \\citet{weak_LLR}. A list of acronyms commonly-used in this paper appear in Appendix~\\ref{app:acronyms}. ", "conclusions": "" }, "0710/0710.0632_arXiv.txt": { "abstract": "We present environmental dependence of the build-up of the colour-magnitude relation (CMR) at $z \\sim 0.8$. It is well established that massive early-type galaxies exhibit a tight CMR in clusters up to at least $z \\sim 1$. The faint end of the relation, however, has been much less explored especially at high redshifts primarily due to limited depths of the data. Some recent papers have reported a deficit of the faint red galaxies on the CMR at $0.8 \\lsim z \\lsim 1$, but this has not been well confirmed yet and is still controversial. Using a deep, multi-colour, panoramic imaging data set of the distant cluster RXJ1716.4+6708 at $z=0.81$, newly taken with the Prime Focus Camera (Suprime-Cam) on the Subaru Telescope, we carry out an analysis of faint red galaxies with a care for incompleteness. We find that there is a sharp decline in the number of red galaxies toward the faint end of the CMR below $M^*+2$. We compare our result with those for other clusters at $z \\sim 0.8$ taken from the literature, which show or do not show the deficit. We suggest that the \"deficit\" of faint red galaxies is dependent on the richness or mass of the clusters, in the sense that poorer systems show stronger deficits. This indicates that the evolutionary stage of less massive galaxies depends critically on environment. ", "introduction": "\\label{sec:intro} It is well known that red early-type galaxies exhibit a tight sequence on colour-magnitude diagrams, which is called the colour-magnitude relation (CMR) (e.g., Visvanathan \\& Sandage 1977; \\citealp{bow92}). In nearby clusters, the CMRs extend down to at least $5-6$ magnitude fainter than the brightest cluster galaxies (e.g., Terlevich, Caldwell \\& Bower 2001). The small colour scatter around the CMR is indicative of the homogeneity of early-type galaxies in clusters (e.g., Bower et al.\\ 1992, 1998). At high redshifts, the CMR has already been well established in clusters at least out to $z\\sim1$ as far as the bright end is concerned (e.g., \\citealt{ell97}; \\citealt{kod98}; \\citealt{sta98}; \\citealp{van98}; 2001; Blakeslee et al.\\ 2003; Stanford et al.\\ 2006; Mei et al.\\ 2006a,b). The faint end of the CMR, however, has been much less explored and still highly uncertain. Some recent deep studies of distant galaxy clusters have shown a relatively small number of galaxies at the faint end of the CMR compared to local clusters. De Lucia et al.\\ (2004, 2007) showed such a deficit of faint red galaxies in $z=0.6-0.8$ clusters observed by the ESO Distant Cluster Survey (EDisCS; \\citealt{whi05}). A similar result was shown in Stott et al.\\ (2007). They compared the faint end of the luminosity functions of red galaxies in $z\\sim 0.5$ clusters from Massive Cluster Survey (MACS; Ebeling et al. 2001) with those of $z \\sim 0.1$ clusters from Las Campanas/AAT Rich Cluster Survey (LARCS; Pimbblet et al. 2006). \\cite{tan05} analysed the RXJ0152.7--1357 cluster (hereafter RXJ0152) at $z=0.83$ based on wide field data taken with the Subaru Prime Focus Camera on the Subaru Telescope (Suprime-Cam; Miyazaki et al.\\ 2002), and they also showed a deficit of faint red galaxies on the CMR. Based on these results, De Lucia et al. (2004, 2007) and \\cite{tan05} discussed that the faint end of the CMR well visible in the present-day universe was established at relatively later epochs as faint blue galaxies stopped their star formation after $z\\sim0.8$ in contrast to much earlier ($z\\gg1$) termination of star formation in massive galaxies. \\cite{tan05} classified galaxy environment into ``cluster'', ``group'' and ``field'', and examined the environmental dependence of the faint end of the CMR as well. They suggest that the build-up of the CMR depends also on environment in the sense that it is more delayed in lower-density environment. The deficit of the faint end of the CMR is often discussed in the line of a currently favoured observational phenomenon called ``down-sizing''. This trend was first noted for field galaxies by Cowie et al.\\ (1996). They showed in their Hawaii Deep Field that most massive galaxies tend to show low star formation rates while less massive galaxies still show on-going star formation activity at $z\\lsim1$. Such a trend has been extended in both redshift space and magnitude range. Kauffmann et al.\\ (2003) showed in the local SDSS data that massive galaxies are dominated by red old galaxies. By contrast, less massive galaxies show bluer colours due to some on-going star formation, and galaxies below a few times $10^{10}$ \\msun {} in stellar mass are predominantly blue. A very similar trend was reported at $z\\sim1$ by Kodama et al.\\ (2004). They looked in the Subaru/XMM Deep Field and showed the distribution of galaxies at $z\\sim1$ on the colour-magnitude diagram. A clear bi-modality on the colour-magnitude diagram was observed again. Since then, a large number of papers have discussed this interesting issue. One of the most convincing cases is based on $\\sim$8,000 galaxies with spectroscopic redshifts within $0.7$2$\\times$10$^{38}$ erg s$^{-1}$ to be proportional to SFR$^{1.06\\pm0.07}$. Furthermore they discovered a linear relation between the total HMXB X-ray flux of a galaxy and its SFR, for SFRs $\\ga$4 M$_{\\odot}$ yr$^{-1}$. Several different methods were used to convert from X-ray intensity to luminosity when creating the XLFs for galaxies in the G03 sample. Some XLFs were derived assuming a standard X-ray binary emission model (a power law with spectral index, $\\Gamma$, $\\sim$1.7 or a 5 keV bremsstrahlung) with Galactic line-of sight absorption (e.g. Zezas et al., 2002; Soria \\& Kong, 2002). Others used the X-ray colours to estimate the emission spectrum, using Galactic absorption (e.g. Eracleous et al., 2002), or deriving the absorption from the colours also (e.g. Lira et al., 2002). Also, some XLFs were derived using best fit spectra to individual bright sources (Smith \\& Wilson, 2001), or to the stacked X-ray population (Roberts et al., 2002). XMM-Newton (XMM) is the most sensitive X-ray imaging telescope in the 0.3--10 keV band. We can therefore use deep XMM observations of nearby galaxies to glean the accuracy of these methods by comparing their resulting XLFs with the XLF derived from freely fitting each X-ray source. To do this, we chose XMM-Newton observations of NGC 253. NGC 253 is a star-bursting spiral galaxy in the Sculptor Group at a distance of $\\sim$4 Mpc; it is almost edge on, with a D$_{25}$ region of $\\sim$25$'$$\\times$7$'$. We fully describe our analysis of the observations in Barnard et al. (2007, MNRAS submitted, hereafter known as B07), and concentrate here on the 2003, 110 ks XMM observation of NGC253. ", "conclusions": "Grimm et al. (2003) report a Universal HMXB XLF derived from published XLFs of several nearby galaxies. They also derive relations between the star formation rate and (i) the total luminosity of the point X-ray sources in the galaxies and (ii) the number of X-ray sources in a galaxy with 2--10 keV luminosity $>$2$\\times$10$^{38}$ erg s$^{-1}$. However, the published XLFs were produced using a number of methods, in most cases assuming Galactic line-of-sight absorption. We have tested several of these models using a deep XMM-Newton observation of the nearby galaxy NGC253, included in a secondary sample of Grimm et al. (2003). We obtained freely modelled luminosities for the 140 brightest sources in the field and also obtained the conversion factors from intensity to flux for some of these different models. We found them to vary by a factor of $\\sim$3. We found the biggest influence on the conversion factor to be the absorption; Model I assumed Galactic line-of-sight absorption, while absorptions 20--50 times higher than this were obtained for the other methods. Since the universal XLF and relations between SFR and X-ray properties were obtained using a mixture of methods, we find them to be inconclusive. We have also found from freely fitting the spectra of 140 bright sources that the corresponence between count-rate and flux is non-linear; this is largely due to the systematic softening of the spectra of more luminous sources. Hence it is unwise to employ a single emission model when describing the X-ray populations of nearby galaxies. Ideally, one would construct XLFs only from luminosities derived from free spectral modelling. To get the most out of the low-photon data, we recommend the stacking method of e.g. Roberts et al. (2002)." }, "0710/0710.4588_arXiv.txt": { "abstract": "{Stability properties of magnetic-field configurations containing the toroidal and axial field are considered. The stability is treated by making use of linear analysis. It is shown that the conditions required for the onset of instability are essentially different from those given by the necessary condition $d (s B_{\\varphi})/ds > 0$, where $s$ is the cylindrical radius. The growth rate of instability is calculated for a wide range of the parameters. We argue that the instability can operate in two different regimes depending on the strength of the axial field and the profile of the toroidal field. ", "introduction": "Turbulence generated by MHD instabilities can play an important role in enhancing transport processes in various astrophysical bodies, such as accretion and protoplanetary disks, galaxies, stellar radiative zones, etc. The anomalous turbulent transport can be particularly important in magnetized gas where a wide variety of MHD instabilities can occur (see, e.g., Barnes et al. 1999). In this case, the onset of instability can be caused both by hydrodynamic motions (for instance, differential rotation; see, e.g., Velikhov 1959; Chandrasekhar 1960) or unstable magnetic configurations. { Which field strength and topology can sustain a stable magnetic configuration is still rather uncertain despite extensive work (see Borra et al. 1982; Mestel 1999 for review).} Most likely, the best-studied magnetic configuration is one with a purely toroidal field. Ever since the paper by Tayler (1973), it has been known that toroidal fields can be unstable close to the axis of symmetry, if there is a non-zero electric current density on the axis. The growth rate of this instability is expected to be of the order of the time taken for an Alfv\\'en wave to travel around the star on a toroidal field line. However, even a purely toroidal field is stable if it decreases rapidly with the cylindrical radius $s$. For instance, Tayler (1973; see also Chanmugam 1979) argued that the toroidal field $B_{\\varphi}$ is stable against axisymmetric perturbations if it satisfies the condition $d ( B_{\\varphi}/s)/ds < 0$ and to non-axisymmetric perturbations if $d (s B_{\\varphi}^2)/ds <0$. Note that a purely toroidal field can also be subject to the magnetic buoyancy instability (Parker 1955; Gilman 1970; Acheson 1978) but the Tayler instability likely appears first as the strength of the toroidal field increases (Spruit 1999). The stability of a purely toroidal field in the radiative zones of stars and accretion disks has been studied by a number of authors. Numerical modeling by Braithwaite (2006) confirms that the toroidal field with $B_{\\varphi} \\propto s$ or $\\propto s^2$ is unstable to the $m=1$ mode ($m$ is the azimuthal wave number) as predicted by Tayler (1973). The linear stability of the toroidal field in rotating stellar interiors has been considered by Kitchatinov \\& R\\\"udiger (2007), who conclude that the magnetic instability is essentially three-dimensional and that the finite thermal conductivity has a strong destabilizing effect. Terquem \\& Papaloizou (1996) and Papaloizou \\& Terquem (1997) considered the stability of an accretion disk with an embedded toroidal magnetic field. These authors find that the disks containing a purely toroidal field are always unstable and obtained spectra of unstable modes in the local approximation. They argue that one type of modes is driven primarily by buoyancy, while the other is driven by shear independently of the magnetic configuration. { Stability properties of purely poloidal magnetic configurations have also been well-studied. Since the papers by Wright (1973) and Markey \\& Tayler (1973, 1974), it is known that a poloidal field is subject to dynamical instabilities in the neighbourhood of neutral points/lines if the field lines are closed inside the star. These authors recognizes that the magnetic field in the neighbourhood of a neutral line resembles that of a toroidal pinched discharge that is known to be unstable. Although instabilities involving significant displacements in the direction of gravity were strongy inhibited, other instabilities were not affected. The instability of poloidal configurations is rather fast: its growth time can reach a few Alfv\\'en crossing time (Van Assche et al. 1982; Braithwaite \\& Spruit 2006). With numerical simulations, the stability of poloidal magnetic configurations has been studied by Braithwaite \\& Spruit (2006), who apply the results to the internal magnetic configuration of neutron stars. Note, however, that a toroidal field might exert a stabilising influence on the instabilities of a poloidal field in the neighbourhood of neutral points (Tayler 1980).} { On the contrary, the addition of even a relatively weak poloidal field alters the stability of the toroidal field substantially. For example, as first shown by Howard \\& Gupta (1962; see also Knobloch 1992; Dubrulle \\& Knobloch 1993), a necessary (but not sufficient) condition for the instability of a toroidal field in the presence of the axial field reads} \\begin{equation}\\label{1} \\frac{d}{ds} (s B_{\\varphi}) > 0. \\end{equation} Howard \\& Gupta (1962) argue that, for a fixed value of $m$, the growth rate of instability caused by condition (1) must vanish in the limit of a vanishing axial magnetic field, thereby providing a connection with the stability criterion obtained by Tayler (1973). Note that the presence of a radial field is also crucial for stability properties of rotating magnetic configurations (Bonanno \\& Urpin 2006). { It turns out that configurations containing both toroidal and poloidal fields are more stable than purely toroidal or purely poloidal ones (Prendergast 1956; Tayler 1980). With numerical simulations, Braithwaite \\& Nordlund (2006) studied the stability of a random initial field in the stellar radiative zone. The star was modeled on a Cartesian grid, and the authors found that the stable magnetic configurations generally have the form of tori with comparable poloidal and toroidal field strengths.} In the present paper, we address the stability properties of magnetic configurations by considering the stability of the toroidal magnetic field with respect to axisymmetric perturbations in the presence of axial fields of a various strength. We show that the instability may occur in such magnetic field configurations under the conditions that differ substantially from those imposed by the Tayler criterion or the necessary condition (\\ref{1}). We argue that, in some cases, the instability is caused by the new type of MHD waves with the growth rate proportional to $\\sqrt{B_z B_{\\varphi}}$ where $B_z$ is the axial magnetic field. Depending on the profile $B_{\\varphi}(s)$ and the ratio $B_z/B_{\\varphi}$, the instability can occur in two regimes that have substantially different growth rates. We also show that the range of unstable wavelength in the $z$-direction can be essentially different, depending on the $B_z/B_{\\varphi}$ ratio. ", "conclusions": "{ We have considered the hydromagnetic stability of cylindrical configurations containing the toroidal and axial magnetic fields. Dissipative effects were neglected in our study. We treated a linear stability assuming that the behaviour of small perturbations is governed by equations of incompressible hydrodynamics. This approximation is justified if the magnetic field is subthermal and the Alfv\\'en velocity is low compared to the sound speed.} The stability of the magnetic configurations is a key issue for understanding the properties of various astrophysical bodies, such as peculiar A and B stars, magnetic white dwarfs, neutron stars, etc. { Even though various dynamo models predict that the toroidal field should be typically stronger than the poloidal one, the effect of a poloidal field on the stability usually cannot be neglected.} To demonstrate this, we treated the simplest model of a highly conducting fluid between two cylindrical surfaces. We assumed that the toroidal and axial fields depends on the cylindrical radius alone. In a short-wavelength approximation, we derived the growth rate and a sufficient criterion of instability analytically (Eqs.~(30)-(31)). For large-scale perturbations, the condition of instability and its growth rate were calculated numerically. The analytical and numerical results are in good qualitative agreement. The obtained conditions of instability differ substantially from what is predicted by the necessary condition (1). For instance, according to Eq.~(1), if the instability occurs in the magnetic configuration, then the toroidal field profile satisfies the condition $\\alpha > -1$. In fact, the instability occurs only if the toroidal field decreases with $s$ much slower (or even increases): the critical value of $\\alpha$ is $\\approx -0.1$ if $B_z/ B_{\\varphi 0} =0.1$ and $\\approx 1$ if $B_z/ B_{\\varphi 0} =0.5$. { If $B_{z}$ depends on $s$, then the critical values of $\\alpha$ should be even higher.} Depending on the profile of the toroidal field and the strength of the axial field, the instability can arise in two essentially different regimes. In the case of a weak axial field, $B_{\\varphi 0} \\gg B_z$, the value of $\\alpha$ that distinguishes between the regimes is $\\approx 1$. If $\\alpha > 1$, then the instability grows on the Alfv\\'en timescale determined by the toroidal field and is rather fast. If $\\alpha < 1$, then the instability is much slower and grows on the timescale determined by the axial field. The transition between these two regimes occurs at larger $\\alpha$ if the axial field increases. The efficiency of the considered instability turns out to be rather low if $\\alpha < 1$ and $B_z$ is weak. It is worth noticing the very particular properties of instability in the case $\\alpha \\approx 1$. In such a configuration, the particular type of MHD waves is given by the dispersion equation (19). The growth rate of these waves (or the frequency, if the waves are stable) is proportional to the product of $B_z$ and $B_{\\varphi}$. These waves cannot exist in purely toroidal or purely poloidal fields. The instability of the configuration with $\\alpha \\approx 1$ is caused by the generation of this particular type of wave. { These waves can probably determine the instability of magnetic configurations near the axis of symmetry where $B_{\\varphi} \\rightarrow 0$.} A sufficiently strong axial field always suppresses the instability. For more or less plausible values of $\\alpha \\leq 1$, the strength of the axial field stabilising the configuration is $\\sim 0.1-1 B_{\\varphi}$. A much stronger field is required, however, to stabilise the configuration with larger $\\alpha$. \\vspace{0.5cm} \\noindent {\\it Acknowledgments.} This research project was supported by a Marie Curie Transfer of Knowledge Fellowship of the European Community's Sixth Framework Programme under contract number MTKD-CT-002995. VU thanks also INAF-Ossevatorio Astrofisico di Catania for hospitality." }, "0710/0710.3682_arXiv.txt": { "abstract": "We report on the latest (2007 Jan) observations of supernova remnant (SNR) 1987A from the {\\it XMM-Newton} mission. Since the 2003 May observations of Haberl et al. (2006), 11 emission lines have experienced increases in flux by factors $\\sim 3$ to 10 ($6 \\pm 0.6$ on average), with the 775 eV line of O~{\\sc viii} showing the greatest increase. Overall, we are able to make Gaussian fits to 17 emission lines in the {\\it RGS} spectra and obtain line fluxes; we have observed 6 lines of Fe~{\\sc xvii} and Fe~{\\sc xviii} previously unreported by {\\it XMM-Newton}. A two-shock model representing plasmas in non-equilibrium ionization is fitted to the {\\it EPIC-pn} spectra, yielding temperatures of $\\sim 0.4$ and $\\sim 3$ keV, as well as elemental abundances for N, O, Ne, Mg, Si, S and Fe. We demonstrate that the abundance ratio of N and O can be constrained to $\\lesssim 20\\%$ accuracy $\\left(\\mbox{N/O} = 1.17 \\pm 0.20 \\right)$. Within the same confidence interval, the same analysis suggests that the C+N+O abundance varies from $\\sim 1.1$ to $1.4 \\times 10^{-4}$, verifying the {\\it Chandra} finding by Zhekov et al. (2006) that the C+N+O abundance is lower by a factor $\\sim 2$ compared to the value obtained in the optical/ultraviolet study by Lundqvist \\& Fransson (1996). Normalizing our obtained abundances by the Large Magellanic Cloud (LMC) values of Hughes, Hayashi \\& Koyama (1998), we find that O, Ne, Mg and Fe are under-abundant, while Si and S are over-abundant, consistent with the findings of Aschenbach (2007). Such a result has implications for both the single-star and binary accretion/merger models for the progenitor of SNR 1987A. In the context of the binary merger scenario proposed by Morris \\& Podsiadlowski (2006, 2007), material forming the inner, equatorial ring was expelled after the merger, implying that either our derived Fe abundance is inconsistent with typical LMC values or that iron is under-abundant at the site of Sanduleak -69$^\\circ$202. ", "introduction": "\\label{sect:intro} SNR 1987A is the Rosetta Stone (Allen 1960) of Type II supernova remnants, resolved and well-studied in multiple wavebands, including the infrared (Bouchet et al. 2006; Kjaer et al. 2007), optical/ultraviolet (Gr\\\"{o}ningsson et al. 2007; Heng 2007) and radio (Gaensler et al. 2007). The physical mechanism partially powering optical/ultraviolet emission from the reverse shock is the same as the one at work in Balmer-dominated supernova remnants (Heng \\& McCray 2007; Heng et al. 2007). The detection of a neutrino burst confirmed the core collapse nature of the progenitor (Koshiba et al. 1987; Svoboda et al. 1987), though a pulsar has yet to be detected (Manchester 2007). A system of three rings may be the result of a binary merger between two massive stars about 20,000 years prior to the supernova explosion (Morris \\& Podsiadlowski 2006, 2007; hereafter MP0607). Reviews of the multi-wavelength studies of SNR 1987A can be found in McCray (1993, 2005, 2007). Mixing of the stellar envelope and core by Rayleigh-Taylor instabilities within the progenitor star, Sanduleak -69$^\\circ$202, has been invoked to explain the early emergence of the 847 keV $\\gamma$-ray line from SNR 1987A, which was predicted by Shibazaki \\& Ebisuzaki (1988) to reach its peak around 1.1 years after the explosion, if one assumes a mixed mass of about 5$M_\\sun$ (Ebisuzaki \\& Shibazaki 1988). Instead, Matz et al. (1988) observed the 847 keV line $\\sim$ 6 to 8 months post-explosion, suggesting even more extensive mixing of $^{56}$Co than assumed. A similar explanation (Ebisuzaki \\& Shibazaki 1988) was given for the early emergence of 16 to 28 keV X-rays (Sunyaev et al. 1990; Inoue et al. 1991). The $\\gamma$-rays originate from the radioactive decay of $^{56}$Co, while the X-rays are from the Compton degradation of the $\\gamma$-rays (McCray, Shull \\& Sutherland 1987). In the soft X-ray, SNR 1987A was first observed by Beuermann, Brandt \\& Pietsch (1994). Subsequently, Hasinger, Aschenbach \\& Tr\\\"{u}mper (1996) tracked a steady increase of the soft X-ray flux over 4 years with {\\it ROSAT}. Extensive work has since been done by the {\\it Chandra} (Michael et al. 2002; Park et al. 2002, 2004, 2005 [hereafter P05], 2006, 2007 [hereafter P07]; Zhekov et al. 2005, 2006 [hereafter Z06]) and {\\it XMM-Newton} groups (Haberl et al. 2006, hereafter H06; Aschenbach 2007, hereafter A07). The general picture gleaned from these studies is of a bimodal plasma distribution present in the region between the forward and reverse shocks (Fig. \\ref{fig:87a}). The soft X-rays ($\\sim$ 0.3 to 0.5 keV) are from the decelerated shock front interacting with dense protrusions (``fingers'') on the inner, equatorial ring, while the ``hard'' X-rays ($\\sim$ 2 to 3 keV) are from a fast shock propagating into more tenuous material. The soft X-rays appear to be correlated with optical ``hot spots'', believed to be emission from the shocked fingers, appearing around the equatorial ring, while the hard X-ray and radio images exhibit structures that coincide. Between Days 6000 and 6200, the soft X-ray light curve experienced an upturn and departure from an exponentially increasing profile, which P05 interpreted as evidence that the blast wave had reached the main body of the dense circumstellar material of the equatorial ring. Future studies on the nature of the soft X-ray light curve are relevant to the issue of pre-ionization of the supernova ejecta, which can potentially extinguish H$\\alpha$ and Ly$\\alpha$ emission from the reverse shock (Smith et al. 2005; Heng et al. 2006). In the studies described, relatively little attention has been paid to the subject of inferring elemental abundances from X-ray analyses. Fits to the X-ray spectra yield N, O, Ne, Mg, S, Si and Fe abundances; such results have been tabulated and studied by Z06 and H06, using {\\it Chandra} and {\\it XMM}, respectively. Here, we examine such an approach using the latest {\\it XMM} data set of SNR 1987A, taken in early 2007. In \\S\\ref{sect:data}, we describe our observations and data reduction techniques. Our results are presented in \\S\\ref{sect:results}. In \\S\\ref{sect:discussion}, we perform a detailed error analysis of individual abundances and their ratios. We find that the N/O ratio as well as the individual N and O abundances can be constrained to $\\sim 20\\%$ accuracy. Our derived elemental abundances for O, Ne, Mg and Fe are under-abundant, while Si and S are over-abundant, relative to typical values for the Large Magellanic Cloud (LMC; Hughes, Hayashi \\& Koyama 1998). Such a result has implications for modeling the progenitor of SNR 1987A, which we discuss. ", "conclusions": "\\label{sect:discussion} \\subsection{INDIVIDUAL ABUNDANCES \\& THEIR RATIOS} \\label{subsect:abund} We compare the derived elemental abundances to those of Z06 and H06 and list them {\\it relative to hydrogen} in Table \\ref{tab:abundances}; uncertainties in the AG89 and W00 abundance tables are not propagated. We emphasize that care must be taken to specify the abundance table used, as this may lead to widely differing values of the derived abundances (relative to hydrogen). In modeling the LMC absorption, Z06 used the elemental abundance table of AG89, while H06 chose the table of W00 because the lower oxygen abundance fitted the K absorption edge in the {\\it EPIC} data better. Fransson et al. (1989) found N/O = 1.6 $\\pm$ 0.8, about 12 times higher\\footnote{The AG89 and W00 values for solar nitrogen-to-oxygen abundance are N/O = 0.132 and 0.110 by number, respectively.} than the AG89 solar value, which they interpret as evidence of substantial CNO processing. LF96 found N/O = 1.1 $\\pm$ 0.4, while Sonneborn et al. (1997) found N/O = 1.7 $\\pm$ 0.5. All three of these values were derived from optical/ultraviolet data. Next, we turn our attention to the N/O ratio derived from X-ray studies. Linearly propagating the errors listed by Z06, we find that their results yield N/O = $1.10^{+0.47}_{-0.45}$; they remark that their derived C, N and O abundance, C+N+O $\\approx 1.98 \\times 10^{-4}$, is lower by about a factor of 2 compared to the $3.72 \\times 10^{-4}$ value of LF96. Model A ({\\tt VNEI+VRAYMOND}) in H06 yields C+N+O $\\approx 1.67 \\times 10^{-4}$ and a rather wide range in the nitrogen-to-oxygen ratio, N/O = $1.33^{+1.47}_{-0.71}$. Our {\\tt VPSHOCK+VPSHOCK} fit to the {\\it EPIC-pn} data yields C+N+O $\\approx 1.29 \\times 10^{-4}$ and N/O = $1.17^{+0.37}_{-0.34}$ (using the W00 table); we call this combination of N and O the ``best fit point''. Note that the C abundance is held fixed at 0.09 relative to solar ($\\sim 3 \\times 10^{-5}$ relative to hydrogen) for the C+N+O values derived from the X-ray studies. We next perform a more careful analysis of the N/O ratio. We first generate a $\\chi^2$ map quantifying the inter-dependence of the fits to the N and O abundance. Contour lines in the $\\chi^2$ map form ``error ellipses'', which are shown for different $\\Delta \\chi^2$ values from the best fit point (Fig. \\ref{fig:NO}). At the $\\Delta \\chi^2 = 2.706$ level, N/O $= 1.17 \\pm 0.20$ for the {\\it EPIC-pn} data. In linearly propagating the errors in the individual abundances, one is in essence adopting the largest possible range of ratios, which can be visualized as the edges of a rectangle in the contour map. Our error analysis improves the uncertainties because it considers only values of the abundance ratio within the specified contour. Within the same confidence interval considered, the corresponding C+N+O value is from $\\sim$ 1.1 to $1.4 \\times 10^{-4}$. We see that the errors in the N/O ratio and the individual N and O abundances can be constrained at the $\\sim 20\\%$ level. We thus confirm the C+N+O under-abundance noted by the {\\it Chandra} studies of Z06, who suggest a couple of physical reasons for such a result: the sub-LMC abundance of C+N+O within the progenitor star, Sanduleak -69$^\\circ$202, and/or an extra source of possibly non-thermal, X-ray continuum. We perform the same analysis for N/S (Fig. \\ref{fig:othercontours}). Again using the W00 table, linear propagation of the errors in the abundances obtained from the {\\tt VPSHOCK+VPSHOCK} fit yields N/S = $4.59^{+1.51}_{-1.36}$, while the error ellipse analysis gives N/S = $4.59^{+1.59}_{-1.33}$. The iron, nitrogen and oxygen lines are predominantly from the lower energy part ($\\lesssim 1$ keV) of the spectrum, while the sulphur lines are situated between $\\sim 2.2$ and 3.1 keV. (The silicon lines are located between $\\sim 1.8$ and 2.2 keV.) Abundance ratios based on lines of widely differing energies lead to rounder error ellipses --- ``error circles''. In such cases, a more thorough error analysis will not constrain the abundance ratio better, as in the case of N/S. This is partially an instrumental effect --- when the energy resolution is comparable to the spacing of the lines, the line complexes overlap and are only partially resolved. There is also the issue of choosing a coordinate system in which the fitting parameters are ``orthogonal''. Hydrogen and helium have no X-ray lines --- their abundances are derived from the strength of the X-ray continuum. When the pair of elements considered are situated close to each other in energy, the hydrogen abundance has to first order a linear dependence on the continuum strength. Thus, the ratio of the considered abundances is tightly constrained as the individual abundances track each other closely. By contrast, when the pair of abundances considered are located far apart in energy, this linear dependence of hydrogen abundance on the continuum is broken as the dominant uncertainties are in temperature rather than in flux. Orthogonality is now absent. To attain orthogonality for such pairs of lines, one has to construct models that directly fit to the abundance ratio considered, an approach which is not explored in this paper. The error in the abundance ratios increases as one moves a given $\\Delta \\chi^2$ away from the minimum point. In Fig. \\ref{fig:ratioerror}, we compute the mean error sustained by the various ratios as a function of $\\Delta \\chi^2$. According to Avni (1976), if three interesting parameters are considered (see \\S\\ref{subsect:spectral}), the 90\\% confidence interval is situated at $\\Delta \\chi^2 = 6.25$. In this case, the N/O abundance ratio suffers from errors $\\sim 25\\%$. \\subsection{THE PROGENITOR OF SNR 1987A: SINGLE-STAR OR BINARY MODEL?} \\label{subsect:metallicity} A more revealing approach to analyze the elemental abundances is to normalize the results listed in Table \\ref{tab:abundances} by the ``canonical'' values of the LMC abundances (Hughes, Hayashi \\& Koyama 1998). These were derived using a sample of 7 middle-aged SNRs in the LMC (N23, N49, N63A, DEM 71, N132D, 0453-68.5 and N49B). This approach was first explored by A07, who showed that the normalized abundances appear to cluster in two groups: N, O, Ne, Mg and Fe are slightly more than half their LMC values, while Si, S and Ni exceed their LMC values. We generalize the A07 approach by considering both sets of abundances derived from using the AG89 and W00 tables. The normalized abundance, relative to its respective LMC value, is $R_{\\rm{87A/LMC}}$; it is plotted as a function of the elemental mass number in Fig. \\ref{fig:abundratios}. The error bars for $R_{\\rm{87A/LMC}}$ are computed by linearly propagating the errors listed in Table \\ref{tab:abundances} and in Hughes, Hayashi \\& Koyama (1998). We caution that additional systematic errors may be present that are not taken into account. For example, the derived abundance for Fe is a sensitive function of temperature and may vary substantially when small changes are made to $T_{\\rm{low}}$ \\footnote{The Fe abundance obtained from the {\\tt VPSHOCK+VPSHOCK} fit (W00 table) varies by $\\sim 16\\%$ when $T_{\\rm{low}}$ is changed by $\\sim 10\\%$.}. We see that the elements O, Ne, Mg and Fe are under-abundant, while Si and S are over-abundant, consistent with the findings of A07. With the exception of Fe, there is a tendency for $R_{\\rm{87A/LMC}}$ to increase with larger elemental mass number, a trend that is independent of the AG89 or W00 tables, though we note that it is more pronounced with the latter. The Fe abundance derived is essentially independent of the AG89 or W00 tables, and is under-abundant by about 70\\% relative to the LMC \\footnote{The O abundance relative to H for the AG89 and W00 tables are the same.}. The under-abundance of Fe and O was previously noted by Hasinger et al. (2006), who argued for the existence of iron-oxygen ``rust grains''. The reduced abundance of Fe alone suggests that the iron is locked up in dust grains. However, Dwek \\& Arendt (2007) showed from an analysis of the infrared-to-X-ray flux ratio --- ${\\cal R}_{\\rm{IRX}} < 1$ versus the theoretically expected value of $\\sim 10^2$ to $10^3$ --- that the dust in SNR 1987A is severely depleted compared to standard dust-to-gas mass ratios in the LMC, suggesting low dust condensation efficiency or dust destruction in the hot X-ray gas. In fact, ${\\cal R}_{\\rm{IRX}}$ was shown to {\\it decrease} with time, which is direct evidence for dust destruction. Our derived plasma temperatures are consistent with this scenario --- even if dust could form, it would be destroyed at these temperatures. In light of Fig. \\ref{fig:abundratios}, the central question to ask is whether the progenitor of SNR 1987A arose from a single star or a binary system? Sanduleak -69$^\\circ$202 was known to be a blue supergiant (BSG) at the time of the supernova explosion, contrary to the expectation that massive stars end their lives as red supergiants (RSGs). Observations of low-velocity, nitrogen-rich circumstellar material are interpreted as the progenitor star being a RSG until about $\\sim 20,000$ years before its death (Fransson et al. 1989). Such a time scale has in turn been interpreted as the Kelvin-Helmholtz time of the helium core (Woosley et al. 1997). This BSG-RSG-BSG evolution remains one of the greatest challenges for the single-star model (Woosley et al. 1997; Woosley, Heger \\& Weaver 2002), as is the observed system of three rings produced $\\sim 20,000$ years before the explosion. The favored single-star models require a combination of reduced metallicity and ``restricted semi-convection'' (Woosley 1988), the former of which is supported by our derived abundance ratios. An additional challenge for the single-star model is to reproduce the over-abundance of Si and S (Fig. \\ref{fig:abundratios}), which may require the invoking of some non-standard mixing process (H.-T. Janka 2007, private communication). Binary solutions to Sanduleak -69$^\\circ$202 are sub-divided into accretion and merger models (see Woosley, Heger \\& Weaver [2002] and references therein). The binary accretion models allow helium- and nitrogen-rich material to be added to the progenitor star and require the disappearance of the mass donor in an earlier supernova event. In the binary merger scenario proposed by MP0607, two stars with masses $\\sim 5 M_\\sun$ and $\\sim 15 M_\\sun$ are initially orbiting each other with a period $\\sim 10$ years. The more massive companion transfers mass to the less massive star only after the former has completed helium burning in the core. A common envelope (Paczy\\'{n}ski 1976) is formed, during which core material from the primary star is dredged up to the surface. The merger process takes a few hundred years, culminating in an initially over-sized RSG, which loses its excess thermal energy over a few thousand years to become a BSG. The spun-up, rapidly rotating BSG produces a fast stellar wind, sweeping up ejecta associated with the merger, producing the triple ring nebula we now see in projection. The nearly axi-symmetric but highly non-spherical nature of the rings suggests that rotation played a role in their formation and is consistent with the proposed scenario. The beauty of the model lies in the fact that it requires no physically ad hoc assumptions --- apart from a small kick of $\\sim 2$ km s$^{-1}$ given to the ejecta to displace the center of the outer rings from their symmetry axis --- and makes a number of predictions. In their favored model, MP0607 assert that the outer rings are ejected before the stellar core material is dredged up, while the inner, equatorial ring --- the site of the observed X-ray emission --- is ejected {\\it afterwards} \\footnote{As noted by MP0607, this hypothesis is verifiable/refutable, as the inner ring should exhibit helium enhancement and more CNO processing relative to the outer rings.}. A relevant consequence of this model is that the dredged-up heavy elements will manifest themselves in the form of X-ray emission lines. This may explain the trend we see in Fig. \\ref{fig:abundratios} --- the challenge for binary merger models is to reproduce the derived $R_{\\rm{87A/LMC}}$ values. Stellar nucleosynthesis can add and {\\it not} subtract iron --- in the context of the MP0607 binary merger model, we expect $R_{\\rm{87A/LMC}}$(Fe) $\\ge 1$. We are again led to the question: where is the iron? If the Fe abundance derived is an upper limit on the iron used to form Sanduleak -69$^\\circ$202, then it is clearly inconsistent with ``standard'' Fe abundances in the LMC. An alternative interpretation is that there is a strong spatial variation in the Fe abundance throughout the LMC, such that the iron is sub-LMC at the site of SN 1987A and equal to its LMC abundance elsewhere. The C+N+O under-abundance suggested in \\S\\ref{subsect:abund} supports such a conclusion. An improvement over using the Hughes, Hayashi \\& Koyama (1998) {\\it ASCA} abundance values is to re-analyze and expand upon their SNR sample using {\\it Chandra} and {\\it XMM}. Existing studies (e.g., Hughes et al. 2006) tend to pick out regions of interest that may include ejecta enrichment; instead, X-ray emission should be extracted from the {\\it entire} blast wave, from which average abundances can be inferred. Future studies will be invaluable towards resolving these issues. \\scriptsize K.H. acknowledges the kind hospitality and financial support of: the Max Planck Institutes for Astrophysics (MPA) and Extraterrestrial Physics (MPE) during June to October 2007, where he was a visiting postdoctoral scientist; and the Lorentz Center (Leiden) during the August 2007 workshop, ``From Massive Stars to Supernova Remnants''. He thanks Dick McCray, Sangwook Park, Svet Zhekov, Jack Hughes, John Raymond, Philipp Podsiadlowski, Rashid Sunyaev, Peter Lundqvist, Claes Fransson, Dmitrijs Docenko, Thomas Janka, Carlos Badenes, Roger Chevalier, Nathan Smith and Jacco Vink for engaging and helpful discussions. Special mentions go out to: Dick and John, who provided a crash course on non-equilibrium ionization plasmas during a sunny bicycle ride in Leiden; Svet, who pointed out relevant material on X-ray physics, as well as for critical comments following his meticulous scrutiny of the manuscript. The authors are collectively grateful to Kazik Borkowski for providing the updated {\\tt VPSHOCK} model files for use in {\\tt XSPEC}. The {\\it XMM-Newton} project is supported by the {\\it Bundesministerium f\\\"{u}r Wirtschaft und Technologie/Deutsches Zentrum f\\\"{u}r Luft- und Raumfahrt} (BMWI/DLR, FKZ 50 OX 001) and the Max Planck Society. \\normalsize" }, "0710/0710.4294_arXiv.txt": { "abstract": "Infrared surveys indicate that the dust content in debris disks gradually declines with stellar age. We simulated the long-term collisional depletion of debris disks around solar-type (G2~V) stars with our collisional code. The numerical results were supplemented by, and interpreted through, a new analytic model. General scaling rules for the disk evolution are suggested. The timescale of the collisional evolution is inversely proportional to the initial disk mass and scales with radial distance as $r^{4.3}$ and with eccentricities of planetesimals as $e^{-2.3}$. Further, we show that at actual ages of debris disks between 10~Myr and 10~Gyr, the decay laws of the dust mass and the total disk mass are different. The reason is that the collisional lifetime of planetesimals is size-dependent. At any moment, there exists a transitional size, which separates larger objects that still retain the ``primordial'' size distribution set in the growth phase from smaller objects whose size distribution is already set by disruptive collisions. The dust mass and its decay rate evolve as that transition affects objects of ever-larger sizes. Under standard assumptions, the dust mass, fractional luminosity, and thermal fluxes all decrease as $t^\\xi$ with $\\xi = -0.3$...$-0.4$. Specific decay laws of the total disk mass and the dust mass, including the value of $\\xi$, largely depend on a few model parameters, such as the critical fragmentation energy as a function of size, the primordial size distribution of largest planetesimals, as well as the characteristic eccentricity and inclination of their orbits. With standard material prescriptions and a distribution of disk masses and extents, a synthetic population of disks generated with our analytic model agrees quite well with the observed Spitzer/MIPS statistics of 24 and 70 \\micron\\ fluxes and colors versus age. ", "introduction": "Since the IRAS discovery of the excess infrared emission around Vega by \\citet{aumann-et-al-1984}, subsequent infrared surveys with ISO, Spitzer and other instruments have shown the Vega phenomenon to be common for main-sequence stars. The observed excess is attributed to second-generation circumstellar dust, produced in a collisional cascade from planetesimals and comets down to smallest grains that are blown away by the stellar radiation. While the bulk of such a debris disk's mass is hidden in invisible parent bodies, the observed luminosity is dominated by small particles at dust sizes. Hence the studies of dust emission offer a natural tool to gain insight into the properties of planetesimal populations as well as planets that may shape them and, ultimately, into the evolutionary history of circumstellar planetary systems. In recent years, various photometric surveys of hundreds of nearby stars have been conducted with the Spitzer Space Telescope. These are the GTO survey of FGK stars \\citep{beichman-et-al-2005,bryden-et-al-2006,beichman-et-al-2006b}, the FEPS Legacy project \\citep{meyer-et-al-2004,kim-et-al-2005}, the A star GTO programs \\citep{rieke-et-al-2005,su-et-al-2006}, the young cluster programs \\citep{gorlova-et-al-2006}, and others. These observations were done mostly at 24 and 70~\\textmu{m} with the MIPS photometer, but also between 5 and 40~\\textmu{m} with the IRS spectrometer \\citep{jura-et-al-2004,chen-et-al-2006}. Based on these studies, about 15\\% of mature solar-type (F0--K0) stars have been found to harbor cold debris disks at 70~\\textmu{m}. For cooler stars, the fraction drops to 0\\%--4\\% \\citep{beichman-et-al-2006b}. For earlier spectral types, the proportion increases to about 33\\% \\citep{su-et-al-2006}. At 24~\\textmu{m}, the fraction of systems with detected excess stays similar for A~stars, but appreciably decreases for FGK ones. Similar results in the sub-millimeter range are expected to become available soon from a survey with SCUBA and SCUBA2 on JCMT \\citep{matthews-et-al-2007}. Preliminary SCUBA results for M dwarfs suggest, in particular, that the proportion of debris disks might actually be higher than suggested by Spitzer \\citep{lestrade-et-al-2006}. All authors point out a decay of the observed infrared excesses with systems' age. However, the values reported for the slope of the decay, assuming a power-law dependence $t^{-\\alpha}$, span a wide range. \\citet{greaves-wyatt-2003} suggest $\\alpha \\la 0.5$, \\citet{liu-et-al-2004} give $0.5 < \\alpha < 1.0$, \\citet{spangler-et-al-2001} report $\\alpha \\approx 1.8$, and \\citet{greaves-2005} and \\citet{moor-et-al-2006} derive $\\alpha \\approx 1.0$. Fits of the upper envelope of the distribution of luminosities over the age yield $\\alpha \\approx 1.0$ as well \\citep{rieke-et-al-2005}. Besides, the dust fractional luminosity exhibits a large dispersion at any given age. In an attempt to gain theoretical understanding of the observed evolution, \\citet{dominik-decin-2003} assumed that equally-sized ``comets'' produce dust through a cascade of subsequent collisions among ever-smaller objects. If this dust is removed by the same mechanism, the steady-state amount of dust in such a system is proportional to the number of comets. This results in an $M/M_0 \\approx \\tau/t$ dependence for the amount of dust and for the number of comets or the total mass of the disk. Under the assumption of a steady state, this result is valid even for more complex systems with continuous size distributions from planetesimals to dust. Tenuous disks, where the lifetime of dust grains is not limited by collisions but by transport processes like the Poynting-Robertson drag \\citep{artymowicz-1997,krivov-et-al-2000,wyatt-2005}, follow $M \\propto t^{-2}$ rather than $M \\propto t^{-1}$. More recently, \\citet{wyatt-et-al-2007a} lifted the most severe simplifying assumption of the Dominik-Decin model, that of equal-sized parent bodies, and included them into the collisional cascade. A debris disk they consider is no longer a two-component system ``comets + dust''. Instead, it is a population of solids with a continuous size distribution, from planetesimals down to dust. A key parameter of the description by \\citet{dominik-decin-2003} is the collisional lifetime of comets, $\\tau$. \\citet{wyatt-et-al-2007a} replaced it with the lifetime of the largest planetesimals and worked out the dependencies on this parameter in great detail. Since the collisional timescale is inversely proportional to the amount of material, $\\tau \\propto 1/M_0$, the asymptotic disk mass becomes independent of its initial mass. Only dynamical quantities, i.e. the disk's radial position and extent, the orbiting objects' eccentricities and inclinations, and material properties, i.e. the critical specific energy and the disruption threshold, as well as the type of the central star determine the very-long-term evolution. Still, there are two important simplifications made in the model by \\citet{wyatt-et-al-2007a}: (i) the disk is assumed to be in collisional equilibrium at all sizes, from dust up to the largest planetesimals and (ii) the minimum specific energy needed to disrupt colliding objects is independent of their size. As a consequence of (i) and (ii), the size distribution of solids is a single power-law. To check how reasonable these assumptions are, realistic simulations of the disks with collisional codes are necessary \\citep[e.g.,][]{thebault-et-al-2003,krivov-et-al-2005,krivov-et-al-2006,thebault-augereau-2007}. The aim of this paper is two-fold. First, we follow the evolution of debris disks with our elaborate numerical code \\citep{krivov-et-al-2005,krivov-et-al-2006} to check the existing analytic models and the assumptions (i) and (ii) they are based upon. Second, in order to make these numerical results easier to use, we develop a new analytic model for the evolution of disk mass and dust mass that relaxes both assumptions (i) and (ii) above. Section~\\ref{secNumerics} summarizes the basic ideas and assumptions and describes our numerical model and the runs of the collisional code. In Section~\\ref{secScalings} the numerical results are presented and dependences of the collisional timescale on the disk mass, distance to the star, and mean eccentricity of parent bodies are derived. In section~\\ref{secAnalytics}, the analytic model for the evolution of disk mass and dust mass is developed. Section~\\ref{secEvolLuminosity} analyzes the evolution of dust luminosities. In Section~\\ref{secObservations}, we use the analytic model to synthesize representative populations of debris disks and compare them with statistics of debris disks derived from the Spitzer surveys. A summary is given and conclusions are drawn in Section~\\ref{secConclusions}. \\pagebreak ", "conclusions": " \\begin{enumerate} \\item The timescale of the collisional evolution is inversely proportional to the initial disk mass. For example, halving the total mass doubles all collisional timescales. This rule is valid for systems where collisions are the only loss mechanism of particles and only as long as $\\beta$-meteoroids are unimportant for the collisional budget. \\item Numerics and analytics consistently yield a $\\tau\\propto r^{4.3}$ dependence of the timescale of the collisional evolution on the radial distance. \\item Numerical simulations show that the collisional timescale varies with the average eccentricity of dust parent bodies as $\\tau \\propto e^{-2.3}$. The analytic approach suggests a somewhat weaker dependence, $\\tau\\propto e^{-5/3}$. \\item An evolving three-slope size distribution is proposed to approximate the numerical results. The biggest objects are still distributed primordially, with a slope $q\\sbs{p}$. The objects below a certain transitional size are already reprocessed by collisions and thus have a quasi-steady-state size distribution, determined by their self-gravity (for intermediate-sized objects, slope $q\\sbs{g}$) or by material strength (for smallest objects, slope $q\\sbs{s}$). That transitional size corresponds to the largest objects for which the collisional lifetime is still shorter than the age of the system. The transitional size increases with time, meaning that ever-larger planetesimals get involved into the collisional cascade. \\item At actual ages of debris disks, $\\sim$10~Myr to $\\sim$10~Gyr, the decay of the dust mass and the total disk mass follow {\\em different} laws. The reason is that, in all conceivable debris disks, the largest planetesimals have longer collisional lifetimes than the system's age, and therefore did not have enough time to reach collisional equilibrium. If the system were let to evolve for sufficiently long time, both dust mass and disk mass would start to follow $t^{-1}$. However, this requires time spans of much longer than 10~Gyr. \\item The loss rate of the dust mass, and the decay rate of fractional luminosity, primarily depend on the difference between the slope $q\\sbs{p}$ of the ``primordial'' size distribution of largest planetesimals and the slope $q\\sbs{g}$ of the size distribution of somewhat smaller, yet gravity-dominated, planetesimals that already underwent sufficient collisional evolution. With ``standard'' values of $q\\sbs{p}$ and $q\\sbs{g}$, the dust mass and the thermal fluxes follow approximately $t^\\xi$ with $\\xi = -0.3\\ldots -0.4$. \\item Specific decay laws of the total disk mass and the dust mass largely depend on a few model parameters. Most important are: the critical fragmentation energy $Q\\sbs{D}^*$ as a function of size, the slope of the ``primordial'' size distribution of planetesimals $q\\sbs{p}$ and their maximum size $s\\sbs{max}$, and the characteristic eccentricity $e$ and inclination $I$ of planetesimals. \\item The property that the maximum possible dust luminosity for a given age does not depend on the initial disk mass, established by \\citet{wyatt-et-al-2007a}, is only valid in cases of very rapid collisional evolution, i.e. in closer-in or dynamically very hot disks. For most of the systems at ages $<10$~Gyr, an increase of the initial disk mass leads to an increase of the dust luminosity, unless that initial mass is assigned extreme values, incompatible with our understanding of planetesimal disks. \\item Assuming standard material prescriptions and disk masses and extents, a synthetic population of disks generated with our analytic model generally agrees with the observed statistics of 24 and 70~\\textmu{m} fluxes versus age. Similarly, the synthetic [24]-[70] colors are consistent with the observed disk colors. \\end{enumerate} As every model, our numerical model makes a number of general simplifying assumptions; the analytic one imposes further simplifications: \\begin{itemize} \\item The collisional evolution is assumed to be smooth and unperturbed. Singular episodes like the aftermath of giant break-ups or special periods of the dynamical evolution such as the late heavy bombardment are not included. \\item Effects of possible perturbing planets are taken into account only indirectly: through the eccentricities of planetesimals (dynamical excitation) and confinement of planetesimal belts (truncation of disks). Further effects such as resonant trapping or ejection of material by planets are neglected. \\item We only consider disruptive collisions. This is a reasonable approximation for disks that are sufficiently ``hot'' dynamically. However, cratering collisions become important when the relative velocities are insufficient for disruption to occur. \\item Neither dilute disks under the regime of Poynting-Robertson drag nor very dense disks with collisional timescales shorter than orbital timescales and with avalanches \\citep{grigorieva-et-al-2007} are covered by the present work. \\item Explaining the initial conditions or deriving them from the dynamical history of the systems at early stages of planetesimal and planetary accretion was out of the scope of this paper. Correlations between disk masses, disk radii, and the presence of planets, for example, were not considered, although they might alter the scalings we found here. \\end{itemize} Despite these limitations, our models reproduce, in essential part, the observed evolution of dust in debris disks. We hope that they may serve as a starting point for in-depth studies that will certainly be undertaken in the future, motivated by questions that remain unanswered, as well as by new data expected from ongoing and planned observational programs." }, "0710/0710.2367_arXiv.txt": { "abstract": "We present measurements of the neutron-capture elements Rb and Pb for bright giants in the globular clusters M4 and M5. The clusters are of similar metallicity ([Fe/H] $\\simeq -1.2)$ but M4 is decidedly $s$-process enriched relative to M5: [Ba/Fe] = +0.6 for M4 but 0.0 for M5. The Rb and Pb abundances were derived by comparing synthetic spectra with high-resolution, high signal-to-noise ratio spectra obtained with MIKE on the Magellan telescope. Abundances of Y, Zr, La, and Eu were also obtained. In M4, the mean abundances from 12 giants are [Rb/Fe] = 0.39 $\\pm$ 0.02 ($\\sigma$ = 0.07), [Rb/Zr] = 0.17 $\\pm$ 0.03 ($\\sigma$ = 0.08), and [Pb/Fe] = 0.30 $\\pm$ 0.02 ($\\sigma$ = 0.07). In M5, the mean abundances from two giants are [Rb/Fe] = 0.00 $\\pm$ 0.05 ($\\sigma$ = 0.06), [Rb/Zr] = 0.08 $\\pm$ 0.08 ($\\sigma$ = 0.11), and [Pb/Fe] = $-$0.35 $\\pm$ 0.02 ($\\sigma$ = 0.04). Within the measurement uncertainties, the abundance ratios [Rb/Fe], [Pb/Fe] and [Rb/X] for X = Y, Zr, La are constant from star-to-star in each cluster and none of these ratios are correlated with O or Na abundances. While M4 has a higher Rb abundance than M5, the ratios [Rb/X] are similar in both clusters indicating that the nature of the $s$-products are very similar for each cluster but the gas from which M4's stars formed had a higher concentration of these products. ", "introduction": "\\label{sec:intro} Globular clusters continue to provide a source of fascination and frustration to both theorists and observers. Two notable accomplishments include the use of globular clusters to (a) check the age of the Universe (e.g., \\citealt{gratton03c}) and to (b) test and refine our understanding of stellar evolution (e.g., \\citealt{renzini88}). Despite these successes, globular clusters continue to present bewildering puzzles. The most persistent puzzle relates to chemical composition. For many years, globular clusters have been known to exhibit star-to-star abundance variations for the light elements C, N, O, Na, Mg, and Al (e.g., see reviews by \\citealt{smith87}, \\citealt{kraft94}, and \\citealt{gratton04}). While the amplitude of the star-to-star abundance dispersion can vary from cluster to cluster, the now familiar anticorrelations between C and N, O and Na, and Mg and Al reveal that the abundance variations are likely produced during hydrogen burning at high temperatures via the CNO, Ne-Na, and Mg-Al cycles. (The O-Na and Mg-Al anticorrelations are not seen in field stars.) However, the stars responsible for the nucleosynthesis and the nature of the pollution mechanism(s) remain poorly understood (see \\citealt{lattanzio06} for a recent summary). One possible explanation for the observed abundance anomalies is internal mixing and nucleosynthesis (e.g., \\citealt{sm79,charbonnel95}) within the present cluster members, the so-called evolutionary scenario. The systematic variation of the C and N \\citep{ss91} and Li \\citep{grundahl02} abundances with luminosity along the red giant branch demand an evolutionary component to the star-to-star abundance variations. Dredge-up of CN-cycled material accounts for the C and N variations. Development of a giant's convective envelope leading to mixing with highly Li-depleted gas accounts for the decline of the Li abundance with increasing luminosity. The proton-capture reactions causing the O, Na, Mg, and Al variations demand much higher temperatures and much deeper mixing than those required for CN-cycling. Such mixing is not predicted by standard theoretical models of red giants and the discovery of the O, Na, Mg, and Al anomalies in main sequence stars (e.g., \\citealt{briley96}, \\citealt{gratton01}) eliminates deep mixing as a viable explanation for the O-Al variations. The interiors of main sequence stars are too cool to process Ne to Na or Mg to Al. Therefore, the cluster gas must have been inhomogeneous when the present stars were formed. This alternative explanation for the abundance anomalies is the so-called primordial scenario. In the primordial scenario, intermediate-mass asymptotic giant branch stars (IM-AGBs) from the generation to which the observed stars belong have long been considered candidates for synthesizing the abundance variations \\citep{cottrell81}. In IM-AGBs, the convective envelope can reach the top of the hydrogen-burning shell, a process called hot-bottom burning. For sufficiently massive and metal-poor AGBs, the temperatures at the base of the convective envelope can exceed 100 million degrees thereby allowing the efficient operation of the CNO, Ne-Na, and Mg-Al cycles (e.g., \\citealt{karakas03}). That IM-AGBs do not alter the abundances of the alpha or iron-peak elements (as required by observations) adds to their qualitative appeal. However, quantitative tests reveal problems with the IM-AGB primordial scenario. Theoretical yields from IM-AGBs combined with a chemical evolution model \\citep{fenner04} suggest that O is not sufficiently depleted, Na is overproduced, Mg is produced rather than destroyed, the isotope ratios of Mg do not match the observations, and the sum of C+N+O increases substantially in contrast to the observations. \\citet{ventura05a,ventura05b,ventura05c} find that many of the flaws noted above can be alleviated when IM-AGB yields are calculated using a revised treatment for convection and mass-loss. However, \\citeauthor{ventura05a} note that problems persist, namely with the Mg isotope ratios, and warn that the predictive power of the current AGB models is limited. Recently, \\citet{prantzos06}, \\citet{smith06}, and \\citet{decressin06} suggest that the winds from massive stars may be more promising candidates than IM-AGBs. There is no satisfactory explanation for the complex patterns for the light element abundances exhibited by every well studied Galactic globular cluster. Therefore, our present understanding of globular cluster chemical evolution and/or stellar nucleosynthesis is incomplete. Determinations of the stellar abundances of the trans-iron or heavy elements offer clues to the history behind the chemical evolution of globular clusters. Here, we provide novel information -- the Rb and Pb abundances -- for giants in M4 and M5, a pair of clusters of similar metallicity but with distinctly different levels of $s$-process products. The quintessential $r$-process element Eu has similar abundances in the two clusters and, indeed, across the collection of Galactic globular clusters. In sharp contrast, the $s$-process products are more evident in M4 than in M5 and other clusters of similar metallicity: [Ba/Fe] is about +0.6 in M4 but 0.0 in M5. The questions - Are there differences in the Rb and Pb abundances between this pair of clusters? and Are the star-to-star variations in the abundances of light elements (O, Na, Mg, and Al) reflected in variations among the abundances of Rb and Pb? -- seem likely to probe the origins of the $s$- and $r$-process products for globular clusters. Due to a critical branching point in the $s$-process path at $^{85}$Kr, the abundance of Rb relative to Sr, Y, or Zr can differ by a factor of 10 depending upon the neutron density at the $s$-process site. In the case of AGB stars, the neutron density in the He-shell is dependent on the stellar mass (e.g., see \\citealt{tomkin83}, \\citealt{lambert95}, \\citealt{busso99}, and \\citealt{abia01} for further details). Since the isotopes of Pb and Bi are the last stable nuclei on the $s$-process path, the $s$-process terminates at these elements and overabundances of Pb and Bi will arise if seed nuclei are shuffled by neutron captures down the entire $s$-process path. In particular, metal-poor AGB stars may produce large overabundances of Pb and Bi if the neutron supply per seed exceeds a critical value (e.g., see \\citealt{goriely01}, \\citealt{travaglio01}, and \\citealt{busso01} for further details). The suspicion is that the star-to-star abundance variations for light elements are due to contamination by IM-AGBs. Some contend that IM-AGBs also synthesize $s$-process nuclides and then one might expect to see star-to-star variations in the Rb and Pb abundances as well as correlations with light element abundances. To further examine the possible role of IM-AGBs in the chemical evolution of globular clusters, \\citet{rbpbsubaru} measured Rb and Pb in NGC 6752 and M13, the two clusters that exhibit the largest amplitude for Al variations. It was found that the abundance ratios [Rb/Zr] and [Pb/Fe] were constant from star-to-star within the measurement uncertainties. If IM-AGBs do synthesize Rb and Pb, then they may not be responsible for the abundance variations. On the other hand, if IM-AGBs are responsible for the abundance variations, they cannot synthesize Rb or Pb. In this paper, we extend the measurements of Rb and Pb to the globular clusters M4 and M5. While these clusters are more metal-rich than NGC 6752 or M13, both M4 and M5 are known to exhibit large dispersions and correlations for the light element abundances [see pioneering studies on CN bimodality by \\citet{norris81a} and \\citet{smith83} as well as recent high-resolution spectroscopic studies by \\citet{M4,M5}, \\citet{ramirez03}, and references therein]. In particular, as noted above, M4 is remarkably, perhaps uniquely among globular clusters, enriched in $s$-process products. ", "conclusions": "\\label{sec:summary} In this paper we present measurements of the neutron-capture elements Rb and Pb in the globular clusters M4 and M5. While both clusters exhibit star-to-star abundance variations for the light elements, we find that the abundances of Rb and Pb are constant. None of the abundance ratios [Rb/Fe], [Rb/Zr], and [Pb/Fe] are correlated with O or Na abundances. In the primordial scenario, the abundance variations for the light elements are attributed to different levels of accretion of ejecta from IM-AGBs or massive stars. The fact that the heavy elements including Rb and Pb do not show abundance variations implies that the accreted material has the same composition as the ambient material for the heavy elements (i.e., the accreted material cannot be highly underabundant or overabundant in these elements). That the ratios [Rb/X] for X = Y, Zr, La are similar for M4 and M5 suggests that the source of the $s$-process elements are similar and that M4 had a greater concentration of these products. There remains a need to pursue additional observational tests of the primordial scenario. In particular, present data on the Rb and Pb abundances in field and cluster stars are sparse. The indication that the Mg isotopic ratios of unpolluted or normal cluster stars differ from those of field stars of the same metallicity deserves closer scrutiny by, in particular, extending the measurement of these isotopic ratios to additional clusters." }, "0710/0710.5637_arXiv.txt": { "abstract": "{ A new method for the determination of open cluster membership based on a cumulative effect is proposed. In the field of a plate the relative $x$ and $y$ coordinate positions of each star with respect to all the other stars are added. The procedure is carried out for two epochs $t_1$ and $t_2$ separately, then one sum is subtracted from another. For a field star the differences in its relative coordinate positions of two epochs will be accumulated. For a cluster star, on the contrary, the changes in relative positions of cluster members at $t_1$ and $t_2$ will be very small. On the histogram of sums the cluster stars will gather to the left of the diagram, while the field stars will form a tail to the right. The procedure allows us to efficiently discriminate one group from another. The greater the distance between $t_1$ and $t_2$ and the more cluster stars present, the greater is the effect. The accumulation method does not require reference stars, determination of centroids and modelling the distribution of field stars, necessary in traditional methods. By the proposed method 240 open clusters have been processed, including stars up to $m<13$. The membership probabilities have been calculated and compared to those obtained by the most commonly used Vasilevskis-Sanders method. The similarity of the results acquired the two different approaches is satisfactory for the majority of clusters. ", "introduction": "Various methods based on the analysis of positions, proper motions, radial velocities, magnitudes and their combinations have been proposed to determine the members of open clusters. The first mathematically rigorous procedure for determination of open cluster membership was developed by Sanders (\\cite{Sanders}) with a statistical analysis of proper motions. It is also the most widely used method. Sanders's approach is based on the model of overlapping distributions of field and cluster stars in the neighborhood and within the region of visible grouping of stars, introduced by Vasilevskis et al. (\\cite{Vasilevskis}). Vasilevskis's model implies that proper motion dispersion of cluster members is caused by observational and measurement errors assumed to be normally distributed. Thus the distribution of cluster stars is represented by a bivariate normal frequency function. The dispersion of field stars is due to not only the errors referred to above, but also to peculiar motion and differential galactic rotation. Therefore the field star distribution is not expected to be random, but rather to have a preferential direction and not normal distribution. However, in a first approximation a bivariate normal ellipsoidal distribution function was assumed for field stars, the major axis of the ellipse being parallel to the galactic plane. Thus, Sanders's equations contain 8 unknown parameters to be determined: the number of cluster members, x and y components of two centroids, one circular and two elliptical dispersions. This system is solved by a maximum likelihood method (Sanders \\cite{Sanders}). The probability of stars being cluster members is calculated by frequency functions with determined parameters. Slovak (\\cite{Slovak}) tested the Vasilevskis-Sanders method by modelling the proper motion distribution in the surroundings of a cluster. He proved the uniqueness and convergence of solutions of the system, provided that errors are represented by a Gaussian distribution and open clusters have no significant internal motion. But if there is noticeable motion within a cluster or it rotates then the above method fails. Cabrera-Ca\\~{n}o and Alfaro (\\cite{Cabrera-Cano Alfaro1}) improved the numerical techniques for obtaining the above parameters. McNamara and Schneeberger (\\cite{McNamara Schneeberger}) showed that the final probabilities could be influenced by various weight groups. Zhao and He (\\cite{Zhao He}) provided a method for treating data with different accuracies. However all these improvements did not treat the problem of star distribution model on which the membership probabilities are based. Even if the hypothesis of a normal distribution of field stars is realistic for some clusters, the centroids of field and cluster stars sometimes are too close to be well discriminated. The parametric model also fails when the cluster member-to-field star ratio is small. The Vasilevskis-Sanders method does not work in the case of significant internal motion in a cluster or its rotation. To overcome some of the problems arising from the parametric Vasilevskis-Sanders method, especially the star distribution modelling, Cabrera-Ca\\~{n}o and Alfaro (\\cite{Cabrera-Cano Alfaro2}) developed a more general, non-parametric method of membership determination. Here no assumptions were made about the nature of cluster and field star distributions and allowed for the use of photometric data. Each cluster needed careful individual study. So, the actual distribution of cluster and field stars, especially the latter, may not be fitted by a Vaselevskis-Sanders model; the centroids for the two groups may be too close to be distinguished; the reliability of the method depends on the cluster-to-field star ratio; in the case of significant internal motions or rotation of a cluster the traditional method fails. In the next section we introduce a new method for the discrimination of cluster stars from surrounding field stars based on a cumulative effect using positions and proper motions. We do not try to overcome all the difficulties of traditional methods, though we offer another approach that enlarges the statistical distance between the two populations - cluster members and field stars - by revealing the group of stars with the least relative velocities. The advantage of this method is that no assumption is made about the distribution of field stars and determination of centroids is avoided. However, in order to determine probabilities we still have to assume a normal bivariate distribution for clusters. Another advantage of the method is that reference stars are not necessary: the discrimination of cluster members is most effective by rectangular coordinates. These features of the method allow us to increase the statistical distance between the two populations. The most noticeable advantage of the accumulation method is its ability to reveal dynamic structures within the clusters if there are any. ", "conclusions": "In the presented method of accumulation we assume that the cluster members moving in the Galaxy as a whole have similar velocities, while field stars show a wide range of velocities. The procedure - adding close velocities while compensating for disperced ones - should enhance the assembling of cluster stars, while distributing more sparsely the field ones, due to the cumulative effect. Thus this method effectively enlarges the statistical distance between physical members and field stars, so that a membership as well as a non-membership is more pronounced. This effect is better for larger cluster-to-field star ratios. For field stars no particular distribution is assumed, thus the centroids are not determined. However, when calculating probabilities, for physical members a normal bivariate distribution function is assumed. For rectangular coordinates no reference stars are needed in the procedure. The most interesting feature of the accumulation mathod is its ability to reveal more then one group of velocities, which has been shown with the example of Pleades. The modified accumulation method was applied to 240 clusters from Dias's list. The probabilities calculated by the accumulation method showed satisfactory agreement with those obtained by the Vasilevskis-Sanders method for the majority of clusters. The poor agreements or disagreements can be ascribed to low cluster-to-field star ratios, or multiple dynamic structures. The accumulation method can be further expanded using other variables like radial velocities and photometric quantities." }, "0710/0710.0224_arXiv.txt": { "abstract": "{ We discuss the large scale properties of standard cold dark matter cosmological models characterizing the main features of the power-spectrum, of the two-point correlation function and of the mass variance. Both the real-space statistics have a very well defined behavior on large enough scales, where their amplitudes become smaller than unity. The correlation function, in the range $0<\\xi(r)<1$, is characterized by a typical length-scale $r_c$, at which $\\xi(r_c)=0$, which is fixed by the physics of the early universe: beyond this scale it becomes negative, going to zero with a tail proportional to $-(r^{-4})$. These anti-correlations represent thus an important observational challenge to verify models in real space. The same length scale $r_c$ characterizes the behavior of the mass variance which decays, for $r>r_c$, as $r^{-4}$, the fastest decay for any mass distribution. The length-scale $r_c$ defines the maximum extension of (positively correlated) structures in these models. These are the features expected for the dark matter field: galaxies, which represent a biased field, however may have differences with respect to these behaviors, which we analyze. We then discuss the detectability of these real space features by considering several estimators of the two-point correlation function. By making tests on numerical simulations we emphasize the important role of finite size effects which should always be controlled for careful measurements. ", "introduction": "In contemporary cosmological models the structures observed today at large scales in the distribution of galaxies in the universe are explained by the dynamical evolution of purely self-gravitating matter (dark matter) from an initial state with low amplitude density fluctuations, the latter strongly constrained by satellite observations of the fluctuations in the temperature of the cosmic microwave background radiation. The other main observational elements for the understanding of the large scale structure of the universe is represented by the studies of galaxy correlations. Any theoretical model aiming to explain the formation of structures must be tested against the data provided by galaxy surveys which give the important bridge between the regimes characterized by large and small fluctuations. Models of the early universe (see e.g. Padmanabhan, 1993 and references therein) predict certain primordial fluctuations in the matter density field, defining the correlations of the initial conditions, i.e. at the time of decoupling between matter and radiation. In the regime where density fluctuations are small enough, the correlation function of the present matter density field is simply related to one describing the initial conditions. In fact, according to the growth of gravitational instabilities in an expanding universe in the linear regime perturbations are simply amplified (see e.g., Peebles, 1980 and references therein). Thus today at some large scales where the correlation function is still positive but with $\\xi(r)<1$ the imprint of primordial fluctuations should be preserved. In the region of strong non-linear fluctuations an analytical treatment to predict the behavior of the two-point correlation function has not been developed yet and, in general, one makes use of numerical simulations which provide a rich, but phenomenological, description of structure in the non-linear regime. It is in this regime, at small enough scales, where most observations have been performed until now. We focus here on the type of correlations predicted in the linear regime by models of the early universe. While the characterization of correlations is usually done in terms of the power-spectrum of the density fluctuations a real space analysis turns out to be useful to point out some relevant features from an observational point of view (see, e.g., the discussion in Gabrielli et al., 2004). Theoretical models of primordial matter density fields in the expanding universe are characterized by a single well-defined length scale, which is an imprint of the physics of the early universe at the time of the decoupling between matter and radiation (see e.g. Bond and Efstathiou 1984, and Padmanabhan 1993 for a general introduction to the problem). The redshift characterizing the decoupling is directly related to the scale at which the change of slope of the power-spectrum of matter density fluctuations $P(k)$ occurs, i.e. it defines the wavenumber $k_c$ at which there is the turnover of the power-spectrum between a regime, at large enough $k$, where it behaves as a negative power-law of the wave number $P(k) \\sim k^{m}$ with $-1 k_c$, and $P(k)\\sim k$ at smaller wavelengths $k r_c$, where the zero-crossing occurs at about $r_c \\approx 124$ Mpc/h in the model considered. We discussed the fact that, globally, a system with this type of correlations belong to the category of super-homogeneous distributions, which are configurations of points more ordered than a purely uncorrelated (Poisson) distribution. Correspondingly fluctuations are depressed with respect to the Poisson case, and the normalized mass variance, for instance, decay faster ($\\sigma^2(r) \\sim r^{-4}$) than for the Poisson case ($\\sigma^2(r) \\sim r^{-3}$). The condition of super-homogeneity is expressed by the condition that $P(k) \\rightarrow 0$ for $k \\rightarrow 0$, or alternatively that \\[ \\int_0^{\\infty} \\xi(r) r^2 dr = 0 \\;. \\] Following the work of Durrer et al. (2003) we have pointed out that the above condition is broken when one samples the distribution, as for example when the simplest biasing scheme of correlated Gaussian fields (introduced by Kaiser, 1984) is applied. This is particularly important for the behavior of the power-spectrum for $kr_c$ is instead expected to be linearly amplified with respect to the original one of the whole matter field. Thus the large scale negative tail $\\xi(r) \\sim -r ^{-4}$ is the main feature which one would like to detect in order to test theoretical models. Given the fact that when $\\xi(r)$ becomes negative, it is characterized by a very small amplitude, the determination of the negative power-law tail represents a very challenging problem. We have discussed the fact that, at first approximation in a real measurement, one may treat the system as having positive correlations at small scales with an exponential cut-off at the scale $r_c$ and then it becomes uncorrelated (a situation which can be regarded as upper limit to the presence of anti-correlations). This implies that for $r_c > 124$ Mpc/h galaxy distribution should not present any positive correlation. Whether this behavior is compatible with the existences of structures of order 200 Mpc/h or more is an open problem which has to be addressed in the studies of forthcoming galaxy catalogs. More in detail, one of the most basic results (see e.g., Peebles 1980) about self-gravitating systems, treated using perturbative approaches to the problem (i.e. the fluid limit), is that the amplitude of small fluctuations grows monotonically in time, in a way which is independent of the scale. This linearized treatment breaks down at any given scale when the relative fluctuation at the same scale becomes of order unity, signaling the onset of the ``non-linear'' phase of gravitational collapse of the mass in regions of the corresponding size. If the initial velocity dispersion of particles is small, non-linear structures start to develop at small scales first and then the evolution becomes ``hierarchical'', i.e., structures build up at successively larger scales. Given the finite time from the initial conditions to the present day, the development of non-linear structures is limited in space, i.e., they can not be more extended than the scale at which the linear approach predicts that the density contrast becomes of order unity at the present time. This scale is fixed by the initial amplitude of fluctuations, constrained by the cosmic microwave background anisotropies (Spergel et al., 2007), by the hypothesized nature of the dominating dark matter component and its correlation properties. According to current models of CDM-type the scales at which non-linear clustering occurs at the present time (of order 10 Mpc) are much smaller than the scale $r_c \\approx 124$ Mpc/h (see e.g. Springel et al., 2005). Thus the region where the super-homogeneous features should still be in the linear regime, allowing a direct test of the initial conditions predicted by early universe models. The scale $r_c$ marks the maximum extension of positively correlated structures: beyond $r_c$ the distribution must be anti-correlated since the beginning, as there was no time to develop other correlations. The possible presence of structures, which mark long-range correlations, whether or not of large amplitude, reported both by observations of galaxy distributions (like the Sloan Great Wall --- see Gott et al., 2005), by the detection of dark matter distributions (see e.g. Massey et al., 2007) and by the large void of radius $\\sim 140$ Mpc identified by Rudnick et al. (2007), is maybe indicating that positive correlations extend well beyond $r_c$. We have discussed that an important finite size effect must be considered when estimating the correlation function, and which may mimic a break of the power-law behavior similar to the ones of CDM models at a scale of order $r_c$. This is related to the effect of the integral constraint in the estimators, namely the fact that the sample average, estimated in a finite sample, differs from the ensemble average, and can be finite-size dependent. This situation occurs when correlations (weak or strong) extend to scales larger than the sample size. For these reasons, in order to study the two-point correlation function in real galaxy samples when its amplitude becomes smaller than unity, it is crucial to check whether the break of the power-law behavior has a finite size dependence or not, by choosing samples with different depth. In this perspective the assessment of the reality of the break of the two-point correlation function is the main observational point to be considered. Once this will be clarified other features should be considered, as for the example the so-called baryonic bump, which is a very small perturbation to the overall shape of the correlation function at scales of order of the zero-point $r_c$. We will present a detailed analysis of the correlation properties of galaxy distribution in the SDSS catalog, considering specific tests for finite-size effects in the determination of the correlation function, in a forthcoming paper." }, "0710/0710.2017_arXiv.txt": { "abstract": "{ H.E.S.S. observations of the old-age ($>$10$^4$~yr; $\\sim 0.5^\\circ$ diameter) composite supernova remnant (SNR) W~28 reveal very high energy (VHE) $\\gamma$-ray emission situated at its northeastern and southern boundaries. The northeastern VHE source (HESS~J1801$-$233) is in an area where W~28 is interacting with a dense molecular cloud, containing OH masers, local radio and X-ray peaks. The southern VHE sources (HESS~J1800$-$240 with components labelled A, B and C) are found in a region occupied by several HII regions, including the ultracompact HII region W~28A2. Our analysis of NANTEN CO data reveals a dense molecular cloud enveloping this southern region, and our reanalysis of EGRET data reveals MeV/GeV emission centred on HESS~J1801$-$233 and the northeastern interaction region. } \\email{growell@physics.adelaide.edu.au} \\begin{document} ", "introduction": "The study of shell-type SNRs at $\\gamma$-ray energies is motivated by the idea that they are the dominant sites of hadronic Galactic cosmic-ray (CR) acceleration to energies approaching the \\emph{knee} ($\\sim 10^{15}$~eV) and beyond, e.g. \\cite{Ginzburg:1}. CRs are then accelerated via the diffusive shock acceleration (DSA) process (eg. \\cite{Bell:1,Blandford:2}). Gamma-ray production from the interaction of these CRs with ambient matter and/or electromagnetic fields is a tracer of such particle acceleration, and establishing the hadronic or electronic nature of the parent CRs in any $\\gamma$-ray source is a key issue. Already, two shell-type SNRs, RX~J1713.7$-$3946 and RX~J0852.0$-$4622, exhibit shell-like morphology in VHE $\\gamma$-rays \\cite{HESS_RXJ1713_II,HESS_VelaJnr_II,HESS_RXJ1713_III} to 20~TeV and above. Although a hadronic origin of the VHE $\\gamma$-ray emission is highly likely in the above cases, an electronic origin is not ruled out. W~28 (G6.4$-$0.1) is a composite or mixed-morphology SNR, with dimensions 50$^\\prime$x45$^\\prime$ and an estimated distance between 1.8 and 3.3~kpc (eg. \\cite{Goudis:1,Lozinskaya:1}). It is an old-age SNR (age 3.5$\\times 10^4$ to 15$\\times 10^4$~yr \\cite{Kaspi:1}), thought to have entered its radiative phase of evolution \\cite{Lozinskaya:1}. The shell-like radio emission \\cite{Long:1,Dubner:1} peaks at the northern and northeastern boundaries where interaction with a molecular cloud \\cite{Wootten:1} is established \\cite{Reach:1,Arikawa:1}. The X-ray emission, which overall is well-explained by a thermal model, peaks in the SNR centre but has local enhancements in the northeastern SNR/molecular cloud interaction region \\cite{Rho:2}. Additional SNRs in the vicinity of W~28 have also been identified: G6.67$-$0.42 and G7.06$-$0.12 \\cite{Yusef:1}. The pulsar PSR~J1801$-$23 with spin-down luminosity $\\dot{E} \\sim 6.2\\times 10^{34}$ erg~s$^{-1}$ and distance $d\\geq9.4$~kpc \\cite{Claussen:3}, is at the northern radio edge. Given its interaction with a molecular cloud, W~28 is an ideal target for VHE observations. This interaction is traced by the high concentration of 1720~MHz OH masers \\cite{Frail:2,Claussen:1,Claussen:2}, and also the location of very high-density ($n>10^3$~cm$^{-3}$) shocked gas \\cite{Arikawa:1,Reach:1}. Previous observations of the W~28 region at VHE energies by the CANGAROO-I telescope revealed no evidence for such emission \\cite{Rowell:1} from this and nearby regions. The High Energy Stereoscopic System (H.E.S.S.: see \\cite{Hinton:1} for details and performance) has observed the W~28 region over the 2004, 2005 and 2006 seasons. After quality selection, a total of $\\sim$42~hr observations were available for analysis. Data were analysed using the moment-based Hillas analysis procedure employing {\\em hard cuts} (image size $>$200~p.e.), the same used in the analysis of the inner Galactic Plane Scan datasets \\cite{HESS_GalScan,HESS_GalScan_II}. An energy threshold of $\\sim 320$~GeV results from this analysis. The VHE $\\gamma$-ray image in Fig.~\\ref{fig:tevskymap} shows that two source of VHE $\\gamma$-ray emission are located at the northeastern and southern boundaries of W~28. The VHE sources are labelled HESS~J1801$-$233 and HESS~J1801$-$240 where the latter can be further subdivided into three components A, B, and C. The excess significances of both sources exceed $\\sim$8$\\sigma$ after integrating events within their fitted, arcminute-scale sizes. Similar results were also obtained using an alternative analysis \\cite{Mathieu:1}. \\begin{figure*}[th] \\centering \\hbox{ \\begin{minipage}{0.55\\textwidth} \\includegraphics[width=\\textwidth]{icrc0129_fig1.eps} \\end{minipage} \\begin{minipage}{0.45\\textwidth} \\caption{H.E.S.S. VHE $\\gamma$-ray excess counts, corrected for exposure and Gaussian smoothed (with 4.2$^\\prime$ std. dev.). Solid green contours represent excess significance levels of 4, 5, and 6$\\sigma$, for an integrating radius $\\theta$=0.1$^\\circ$. The VHE sources HESS~J1801$-$233 and a complex of sources HESS~J1800-240 (A, B \\& C) are indicated. The thin-dashed circle depicts the approximate radio boundary of the SNR W~28 \\cite{Dubner:1,Brogan:1}. Additional objects indicated are: HII regions (black stars); W~28A2, {G6.1$-$0.6} % {6.225$-$0.569}; % The 68\\% and 95\\% location contours (thick-dashed yellow lines) of the $E>100$~MeV EGRET source {GRO~J1801$-$2320}; the pulsar {PSR~J1801$-$23} (white triangle). The inset depicts a pointlike source under identical analysis and smoothing as for the main image.} \\label{fig:tevskymap} \\end{minipage} } \\end{figure*} ", "conclusions": "H.E.S.S. and NANTEN observations reveal VHE emission in the W~28 region spatially coincident with molecular clouds. The VHE emission and molecular clouds are found at the northeastern boundary, and $\\sim 0.5^\\circ$ south of W~28 respectively. The SNR W~28 may be a source of power for the VHE sources, although there are additional potential particle accelerators in the region such as other SNR/SNR-candidates, HII regions and open clusters. Further details concerning these results and discussion are presented in \\cite{HESS_W28}." }, "0710/0710.5727_arXiv.txt": { "abstract": "{}{We communicate the detection of soft (20--200\\,keV) $\\gamma$-rays from the pulsar and pulsar wind nebula of \\hbox{PSR~J1846$-$0258} and aim to identify the component of the system which is responsible for the $\\gamma$-ray emission.}{We combine spectral information from the \\integral $\\gamma$-ray mission with archival data from the \\chandra X-ray Observatory to pinpoint the source of soft $\\gamma$-ray emission.}{Our analysis shows that the soft $\\gamma$-rays detected from \\hbox{PSR~J1846$-$0258} include emission from both the pulsar and the pulsar wind nebula, but the measured spectral shape is dominated by the pulsar wind nebula. We discuss \\hbox{PSR~J1846$-$0258} in the context of rotation powered pulsars with high magnetic field strengths and review the anomalously high fraction of spin-down luminosity converted into X- and $\\gamma$-ray emission in light of a possible overestimate of the distance to this pulsar.}{} ", "introduction": "The pulsar \\hbox{PSR~J1846$-$0258} (also known as \\hbox{AX~J1846.4$-$0258}) was discovered by \\citet{GotthelfVasishtBoylan-Kolchin2000} in the X-ray band and lies near the centre of the supernova remnant Kes 75 (SNR G29.7-0.3). Recent X-ray imaging \\citep{HelfandCollinsGotthelf2003} shows that the pulsar is embedded in a pulsar wind nebula (PWN) which shows distinct physical structure. No radio emission has been observed from \\hbox{PSR~J1846$-$0258}, but X-ray timing reveals the pulse period to be $P=324$\\,ms and the characteristic age is $P/(2\\dot{P})=728$\\,years -- the smallest characteristic age of any rotation-powered pulsar. The inferred surface dipole magnetic field strength is $4.8\\times10^{13}$\\,G -- almost an order of magnitude greater than the magnetic fields for typical pulsars, placing it closer to the field strengths of magnetars. The distance to the system is $\\sim19$\\,kpc as estimated from the neutral hydrogen density along the line of sight \\citep{BeckerHelfand1984}, which implies that the size of the supernova remnant (SNR) is extremely large for such a young pulsar. It also implies that the efficiency of the conversion of the pulsar spin-down luminosity ($8.4\\times10^{36}$\\,erg\\,s$^{-1}$, using $P$ and $\\dot P$ from \\citealt{GotthelfVasishtBoylan-Kolchin2000}) to combined pulsar and PWN X-ray luminosity in the 0.5--10\\,keV energy range, is 20\\% (using the flux from this work) -- the largest of any rotation-powered pulsar. This source is one of a growing class of rotation-powered pulsars with B-field strengths approaching those of magnetars. In this paper we present the \\integral observations and results for \\hbox{PSR~J1846$-$0258}, link these results to previously published \\chandra results and follow up with a discussion of the properties of this source in comparison to other members of this class and examine the possibility of an overestimate in the distance to this pulsar. \\begin{figure*} \\sidecaption \\includegraphics[width=12cm]{8432fig1.eps} \\caption{IBIS/ISGRI significance mosaic of the 20--100\\,keV energy band, showing \\hbox{PSR~J1846$-$0258} in the centre of the field. False colour representation of the significance is displayed on a logarithmic scale. The top right inset is a $50\\arcsec\\times50\\arcsec$ image from our reduction of \\chandra data in the 0.3--10\\,keV energy band. One can clearly distinguish the bright pulsar and the surrounding synchrotron nebulosity. The white ellipses on the \\chandra image indicate the extraction regions for the PWN and pulsar spectra as explained in Sect.~\\ref{chanresults}.} \\label{FigIntimg} \\end{figure*} ", "conclusions": "\\label{discuss} PSR~J1846$-$0258 is one of a recently emerged and growing class of rotation-powered pulsars with inferred surface dipole magnetic fields approaching those of the magnetars. A number of these pulsars have been discovered as a result of the Parkes multi-beam pulsar survey of the Galactic plane \\citep{ManchesterLyneCamilo2001}, for example \\hbox{PSR~J1119$-$6127} and \\hbox{PSR~J1814$-$1744} \\citep{CamiloKaspiLyne2000}, which have inferred magnetic fields of $4.1\\times10^{13}$\\,G and $5.5\\times10^{13}$\\,G respectively. These are all young pulsars with characteristic ages of order a few thousand years and approximately half of them show X-ray emission (\\hbox{PSR~J1119$-$6127}, \\citealt{GonzalezSafi-Harb2003}; \\hbox{PSR~J1718$-$3718}, \\citealt{KaspiMcLaughlin2005}) with spectra harder than those of magnetars. In fact, \\hbox{PSR~J1119$-$6127} shows a remarkable resemblance to \\hbox{PSR~J1846$-$0258} as far as spin characteristics and the properties inferred from them are concerned. \\hbox{PSR~J1119$-$6127} has $P=408$\\,ms and $\\dot{P}=4.1\\times10^{-12}$\\,s\\,s$^{-1}$, giving a spin-down luminosity, characteristic age and inferred surface dipole magnetic field very close to those of \\hbox{PSR~J1846$-$0258} \\citep{CamiloKaspiLyne2000}. This, however, is where the similarity ends. Across the electromagnetic spectrum, these pulsars have very different attributes. In the soft $\\gamma$-ray band covered by IBIS/ISGRI, \\hbox{PSR~J1846$-$0258} is clearly detected in the 20--100\\,keV band (see Table~\\ref{TabSpecPar}), while with similar exposure we can place a $2\\sigma$ upper limit of $3.3\\times10^{33}$\\,erg\\,s$^{-1}$ on any emission from the location of \\hbox{PSR~J1119$-$6127} (using a distance of 8.4\\,kpc, \\citealt{CaswellMcClure-GriffithsCheung2004}) in the same energy band. Also in the X-ray band (0.5--10\\,keV), these two pulsars show contrasting spectral behaviour. While \\hbox{PSR~J1846$-$0258} shows absorbed power law spectra from both the pulsar and PWN \\citep{HelfandCollinsGotthelf2003}, \\hbox{PSR~J1119$-$6127} shows a strong thermal spectrum superimposed on a hard power law \\citep{GonzalezKaspiCamilo2005}. Although \\citet{GonzalezKaspiCamilo2005} attribute the hard power law component to the PWN, the thermal X-rays from the pulsar are difficult to interpret through conventional emission models, e.g. models for initial cooling of the entire neutron star result in a radius smaller than allowed by neutron star equations of state, while hot spot models result in unusually high temperature estimates. It may be the case that, while these two pulsars show similar current spin characteristics, their evolutionary paths and physical properties may be very different, as reflected in the significantly different emission properties. The most unusual feature of \\hbox{PSR~J1846$-$0258} is its efficiency in converting spin-down power, $\\dot E$, into X- and $\\gamma$-ray luminosity $L_{\\rm X}$ and $L_{\\rm \\gamma}$. Using data from 41 pulsars, \\citet{PossentiCeruttiColpi2002} determined a general relation between the 2--10\\,keV X-ray luminosity of the combined pulsar and PWN, and the spin-down power. According to this relation, and the value of $\\dot E$ determined from the spin characteristics, $L_{\\rm 2-10\\,keV}/\\dot{E}$ should be $\\sim 0.2$\\% for \\hbox{PSR~J1846$-$0258}. The measured efficiency, at $L_{\\rm 2-10\\,keV}/\\dot{E}\\sim 12$\\%, when combined with the 20--100\\,keV measured in this work indicates $L_{\\rm 2-10,20-100\\,keV}/\\dot{E}\\sim27$\\%. It is expected that the total efficiency in converting spin-down power into luminosity is actually much larger than this, as the luminosity contribution between 10--20\\,keV is not included and the soft $\\gamma$-ray flux extends beyond 100\\,keV (see Fig.~\\ref{FigSpec}). The observation that the 2--10\\,keV spin-down conversion efficiency is so much larger than average, and the fact that this efficiency becomes even larger upon inclusion of the 20--100\\,keV luminosity, may be a reason to call the distance estimate into question. To this effect, we can compare the efficiency of \\hbox{PSR~J1846$-$0258} with two similar young pulsars in the Large Magellanic Cloud, for which the distance is well known at $\\sim50$\\,kpc. Including luminosity contributions from both the pulsar and PWN \\hbox{PSR~J0537$-$6910} has $L_{\\rm 0.5-10\\,keV}/\\dot{E}=0.9$\\% \\citep{ChenWangGotthelf2006} and \\hbox{PSR~B0540$-$69} has $L_{\\rm 0.5-10\\,keV}/\\dot{E}=9.1$\\% \\citep{KaaretMarshallAldcroft2001}. Even though the estimate for \\hbox{PSR~B0540$-$69} is much larger than the X-ray efficiency of most pulsars, it is still only around half that of \\hbox{PSR~J1846$-$0258} ($L_{\\rm 0.5-10\\,keV}/\\dot{E}=20$\\%). \\citet{BeckerHelfand1984} determined a distance of 19\\,kpc through neutral hydrogen measurements. This distance, which agrees with the estimate of \\citet{Milne1979} based on the surface brightness -- diameter relation for supernovae and the neutral hydrogen measurements of \\citet{CaswellMurrayRoger1975}, does imply a very high mean expansion velocity for the supernova remnant \\citep{HelfandCollinsGotthelf2003}. On the other hand, if the distance were overestimated, this would lead to inconsistency between the neutral hydrogen column as measured by \\citet{BeckerHelfand1984} and that deduced from fits to X-ray data of the PWN \\citep{HelfandCollinsGotthelf2003} -- such inconsistency is not observed. \\citet{MortonSlaneBorkowski2007} also discuss the possibility of an incorrect distance estimate, but dismiss this option as it would double the already high inferred density of postshock gas in the SNR. We can estimate a lower limit on the distance of the system by employing the half width at half maximum of the Si line as measured by \\citet{HelfandCollinsGotthelf2003}. Assuming that the SNR has been expanding at this measured rate (1850\\,km\\,s$^{-1}$) for its entire lifetime (an upper limit of 884 years, \\citealt{LivingstonKaspiGotthelf2006}), we can estimate the distance from the angular size of the SNR on the sky to be $\\sim3$\\,kpc. The real distance is certainly larger than this lower limit due to the fact that the SNR expansion velocity at earlier times was most likely larger than our current estimate. Using this lower limit on the distance would provide $L_{\\rm 0.5-10\\,keV}/\\dot{E}\\sim 0.5$\\% -- more in line with the efficiencies of other rotation-powered pulsar systems. This makes the prospect of a distance smaller than 19\\,kpc to \\hbox{PSR~J1846$-$0258} appealing, but difficult to substantiate in terms of the existing estimates \\citep{CaswellMurrayRoger1975,Milne1979,BeckerHelfand1984}. Of the rotation-powered pulsars that exhibit X-ray emission, only very few have been detected above 10\\,keV. Although a consistent treatment exploring the general properties of PWN in the 20--100\\,keV range will be presented as a future publication, we can see from the existing literature that the \\integral spectrum of \\hbox{PSR~J1617$-$5055}, a young rotation powered pulsar, shows a power law slope of $\\sim 2$ \\citep{LandiDeRosaDean2007}, while a combined IBIS/ISGRI and BeppoSAX spectrum of the PWN in \\hbox{PSR~B1509$-$58} \\citep{ForotHermsenRenaud2006} can be fit with a power law of photon index $\\sim2.1$ -- both close to the spectral shape of \\hbox{PSR~J1846$-$0258}. In addition, the EGRET instrument aboard the Compton Gamma-Ray Observatory detected emission from $\\sim7$ rotation-powered pulsars and/or their PWN in the 30\\,MeV--20\\,GeV energy range \\citep{Roberts2005}. These pulsars have photon indices harder than 2.2, also consistent with the soft \\gm ray spectral slope of \\hbox{PSR~J1846$-$0258} in this work and with those \\integral observations quoted above. However, smooth spectral coverage above 100\\,keV and into the GeV energy range is required to ascertain the presence or absence of spectral turnovers predicted by models of PWN emission mechanisms \\citep{ZhangHarding2000,ChengHoRuderman1986a}." }, "0710/0710.5511_arXiv.txt": { "abstract": "% In hierarchical galaxy formation the stellar halos of galaxies are formed by the accretion of minor satellites and therefore contain valuable information about the (early) assembly process of galaxies. Our GHOSTS survey measures the stellar envelope properties of 14 nearby disk galaxies by imaging their resolved stellar populations with HST/ACS\\&WFPC2. Most of the massive galaxies in the sample ($V_{\\rm rot}$$>$200 km/s) have very extended stellar envelopes with $\\mu(r)$$\\sim$$r^{-2.5}$ power law profiles in the outer regions. For these massive galaxies there is some evidence that the stellar surface density of the profiles correlates with Hubble type and bulge-to-disk ratio, begging the question whether these envelopes are more related to bulges than to a Milky Way-type stellar halo. Smaller galaxies ($V_{\\rm rot}$$\\sim$100 km/s) have much smaller stellar envelopes, but depending on geometry, they could still be more luminous than expected from satellite remnants in hierarchical galaxy formation models. Alternatively, they could be created by disk heating through the bombardment of small dark matter sub-halos. We find that galaxies show varying amounts of halo substructure. ", "introduction": "We select RGB stars from our CMDs and use those to trace the stellar surface density along the minor axis. RGB stars are ideal as they are abundant in our CMDs, are indicative of old stellar populations (as expected to be found in the outskirts of galaxies), and are representative of the underlying stellar mass. To map the surface brightness profiles in the central regions of the galaxies we use the integrated light from Spitzer/IRAC 4.5 micron observations. The Spitzer images provide near unobscured light profiles, even for our edge-on galaxies. We scale the RGB surface density star counts such that they match the IR luminosity profiles in the overlapping region. In this way we derive equivalent surface brightness profiles directly from the RGB star counts. \\begin{figure} \\mbox{ \\epsfclipon \\epsfxsize=0.49\\textwidth \\epsfbox[83 207 500 550]{prof_min_linr00.eps} \\epsfclipon \\epsfxsize=0.49\\textwidth \\epsfbox[83 207 500 550]{prof_min_logr01.eps} } \\vspace*{-5mm} \\caption{ Minor axis surface density profiles of GHOSTS galaxies. The thin solid lines indicate the profiles derived from Spitzer/IRAC 4.5 micron images calibrated to Vega magnitudes (add about 3.5 mag to convert to Vega $V$-mag). The symbols connected with dotted lines represent RGB star count profiles, scaled to match the Spitzer data. To reduce confusion at small radii we only plot star counts beyond 12 kpc for the non-edge-on galaxies NGC\\,3031/M81 and NGC\\,5236/M83. On a linear radial scale (left diagram) exponential disks appear as straight lines, as indicated for IC5052 by the dot-dashed thick line. In the log-log plot on the right, where we have removed low mass galaxies for clarity, a straight line indicates a power law profile (e.g. thin dashed line = r$^{-2.5}$). Also show are an exponential disk (for NGC\\,0891, dot-dashed line) and a S\\'ersic profile with the typical parameters for a flattened stellar halo as modeled by \\citet{AbaNav06} (thick dashed line). \\vspace{-1mm} \\label{profs} } \\end{figure} In Fig.\\,\\ref{profs} we show profiles of the edge-on galaxies analyzed so far, along with outer profiles for the more face-on galaxies NGC\\,3031/M81 and NGC\\,5236/M83% . The exponential thin disks only dominate the inner $\\sim$2--3 kpc (5-10 scale heights), while the extended components are evident at larger radii. We find that eight of the nine galaxies analyzed thus far show components that are more extended than the exponential disks detected at small radii. Several of the most massive galaxies have very extended envelopes with stellar densities at 30 kpc that are 10--100 times higher than the contamination background, equivalent to $\\sim$29 $V$-mag arcsec$^{-2}$. NGC\\,5236/M83 is the only galaxy with a pure exponential disk to the last measured point at 20 kpc (more than 10 disk scale lengths). \\subsubsection{The bulge-halo connection} In this section we explore the connection between bulges and the extended components. We model the minor axis profiles by combining a S\\'ersic profile and an exponential disk. Merger models show that the hot components resulting after a violent relaxation generally exhibit a S\\'ersic profile \\citep[e.g.,][and reference therein]{BarHer92,AbaNav06}. If bulges and stellar envelopes are created by a collisionless merger processes, we thus expect their light to follow a S\\'ersic profile. For a number of massive galaxies ($V_{\\rm rot}$$>$200 km\\,s$^{-1}$) we can fit the entire minor axis profile over a factor of 1000 in size ($\\sim$10$^{4.5}$ in surface density) by an exponential disk and a single S\\'ersic profile, representing both the inner bulge-like region and the outer envelope. This is, for instance, the case for the bulge dominated NGC\\,7814 or a galaxy like NGC\\,3031/M81, which has a power law outer envelope of rather steep slope. Other galaxies, like NGC\\,4565% , have too shallow an outer slope compared to their concentrated bulge region to be fitted by a S\\'ersic profile. NGC\\,891 can be fitted by an exponential disk and a S\\'ersic profile from central bulge to outer envelope if we ignore our outermost field and presume that the higher star density at 30 kpc is due to substructure. Finally NGC\\,5236/M83, which has only a small bulge, shows no sign of an outer envelope out to ten disk scale lengths. The smallest galaxies in the sample ($V_{\\rm rot}$$\\simeq$100--120 km\\,s$^{-1}$) have small extended components, barely discernible above the background contamination (see Fig.\\,\\ref{profs}). The shape of the extended component is thus poorly constrained due to both the uncertainty in the background and low number statistics. The star counts can be fitted equally well by exponential, power law, and S\\'ersic law profiles. This feature could be the thick disk, as observations of the NGC\\,4244 major axis suggest the component is very flattened. However, the RGB main disk scale height is already twice that of the main sequence population and has been argued already represent the thick disk \\citep{Seth05II} with the feature observed here being an additional component. These additional components are most likely (depending on exact shape) more luminous than predicted in the hierarchical models of \\citet{PurBul07}, but could have been created by the bombardment of small dark matter sub-halos \\citep{KazBul07}. Therefore, a number of the observed extended envelopes seem structurally directly related to the central bulge regions, like in NGC\\,7814. In other galaxies, where the bulge is too concentrated to be simply related to the outer envelope, we can suspect that secular evolution (e.g., bar driven central enhancement and thickening) can account for the extra (pseudo-)bulge light. NGC\\,4565 with its boxy bulge could nicely fall in this category. In small galaxies the extended envelope is unrelated to the central region, as these small galaxies have no bulge. \\subsubsection{Envelope properties and halo models} Comparing the envelopes of the different galaxies we find that the two small galaxies have much smaller extended components than the larger galaxies, with surface densities that are lower relative to their disks. The more massive galaxies in our sample are very similar in terms of mass, luminosity, and scale size. Still, there is significant variation in outer envelope properties. At 20 kpc NGC\\,891, NGC\\,4565, and NGC\\,7814 have power law profiles with a slope of about -2.5. NGC\\,3031/M81 has a steeper profile, while at 20 kpc NGC\\,5236 is still dominated by the (face-on) disk. At first sight, the envelope luminosity at 20 kpc seems correlated with Hubble type and bulge-to-disk ratio, with the bulge dominated NGC\\,7814 being the brightest and the late-type spiral NGC\\,5236 showing no sign of an envelope at all. Although M31 also fits this trend, the Sab galaxy NGC\\,3031/M81 does not, as a steeper and fainter profile is evident at 20 kpc. In Fig.\\,\\ref{profs} we also show a typical profile from the \\citet{AbaNav06} model of accreted stars. Our profiles are a bit shallower and mostly fainter than these models between 10 and 30 kpc. While the surface density normalization may be somewhat uncertain in the models, the shape is quite well constrained. It could be that the true halos only dominate at even larger radii and the slope becomes even shallower at larger radii. However, in hierarchical galaxy formation the halo and ``classical'' bulge are formed by the same merging process, so there is no reason to suspect a large structural change between bulge and halo. The S\\'ersic radii derived for our combined envelope and bulge fits are typically eight times smaller than those of \\citet{AbaNav06}. However, a number of simulation parameters affect the concentration of the accreted halos. Increasing star formation suppression in small sub-halos, such that only the most massive dark matter sub-halos contain stars, yields steeper and fainter envelopes \\citep{BekChi05}. Alternatively, the stars in the accreted satellites could sit deeper in the potential wells of their dark matter sub-halos than simulated in these models, thereby only being tidally stripped closer to the main galaxy, also resulting in more concentrated halos \\citep{BulJoh05}. ", "conclusions": "" }, "0710/0710.0272_arXiv.txt": { "abstract": "{ANTARES is a large volume neutrino telescope currently under construction off La Seyne-sur-mer, France, at 2475m depth. Neutrino telescopes aim at detecting neutrinos as a new probe for a sky study at energies greater than 1 TeV. The detection principle relies on the observation, using photomultipliers, of the Cherenkov light emitted by charged leptons induced by neutrino interactions in the surrounding detector medium. Since late January 2007, the ANTARES detector consists of 5 lines, comprising 75 optical detectors each, connected to the shore via a 40 km long undersea cable. The data from these lines not only allow an extensive study of the detector properties but also the reconstruction of downward going cosmic ray muons and the search for the first upward going neutrino induced muons.The operation of these lines follows on from that of the ANTARES instrumentation line, which has provided data for more than a year on the detector stability and the environmental conditions. The full 12 line detector is planned to be fully operational early 2008.} \\begin{document} ", "introduction": " ", "conclusions": "Great achievements have been made by the Antares collaboration in the last year. The detector is working in nominal mode with 5 lines and should be complete early 2008. Upward neutrino candidates have been found that validate the conceptual method and the chosen techniques. Very exciting times have started with a detector looking for neutrinos in a region of the celestial sky which has never been studied with such a level of sensitivity." }, "0710/0710.0758_arXiv.txt": { "abstract": "The star HDE 226868 known as an optical counterpart of the black hole candidate Cyg X-1 has been observed in H$_\\alpha$ region using spectrograph at Ond\\v{r}ejov 2-m telescope. The orbital parameters are determined from He\\,I-line by means of the author's method of Fourier disentangling. Preliminary results are also presented of disentangling the H$_\\alpha$-line into a P-Cyg profile of the (optical) primary and an emission profile of the circumstellar matter (and a telluric component). ", "introduction": "The bright X-ray source Cyg X-1 has been identified with the star denoted as HDE 226868, V1357 Cyg or BD+34$^{\\circ}$3815 etc. An improvement of instrumentation of the Ond\\v{r}ejov 2-m telescope enabled to start with systematic observations of this target of magnitude V$\\simeq$8.9, B$\\simeq$9.6. With coordinates $\\alpha_{2000}=19^{h}58^{m}21.7^{s}$, $\\delta_{2000}=+35^{\\circ}12'6''$ it is well observable from Ond\\v{r}ejov mainly at summer. It is known to be an interacting binary with period $P\\simeq 5.6 d$. The primary component is a supergiant of spectral type classified as B0 (or O9.7) Iab with temperature $T_{\\rm eff}=30400\\pm500$ K and log $g=3.31\\pm0.07$. This primary, which nearly fills its Roche lobe, shows signs of variable strong stellar wind and an overabundace of He and heavier elements (cf. e.g. Karitskaya et al. 2007). The secondary component invisible in optical radiation is a compact object, most probably a black hole. This companion, or its neighborhood emits a variable X-radiation, which is supposed to originate from an accretion disk fed by the stellar wind from the primary. The X-radiation switches chaotically between two states. In the low/hard state the total X-ray flux is low and the spectrum is flat, so that the hard tail of X-radiation prevails. In the high/soft state the soft radiation is enhanced more, and consequently the spectrum has a steeper decrease toward the higher energies and hence the radiation is softer in the mean. Some intermediate states may also appear temporarily. The X-ray flux is anticorrelated with the strength of emission in the H$_{\\alpha}$-line of the primary: in the X-low/hard state the H$_{\\alpha}$ emission is strong, while in the X-high/soft state the H$_{\\alpha}$ emission is weak. The aim of the observational campaign at Ond\\v{r}ejov observatory was to improve orbital parameters of the system, to check a possible spectroscopic features connected with the circumstellar matter (either accretion disk around the black hole, gaseous streams or stellar wind) or with a possible third body, and to get line-profiles enabling a quantitative comparison with a model of the atmosphere and stellar wind of the primary. The first part of obtained spectra was provided for a study on Cyg X-1 organized in a wide international collaboration, the results of which should appear in Gies et al. (2007). In the present contribution, results obtained using the author's method of spectra disentangling from the same set of Ond\\v{r}ejov spectra are given. A more detailed study taking into account also recently obtained spectra is in progress. ", "conclusions": "" }, "0710/0710.0875_arXiv.txt": { "abstract": "Theoretical arguments and indirect observational evidence suggest that the stellar initial mass function (IMF) may evolve with time, such that it is more weighted toward high mass stars at higher redshift. Here we test this idea by comparing the rate of luminosity evolution of massive early-type galaxies in clusters at $0.02\\leq z\\leq 0.83$ to the rate of their color evolution. A combined fit to the rest-frame $U-V$ color evolution and the previously measured evolution of the $M/L_B$ ratio gives $x = -0.3^{+0.4}_{-0.7}$ for the logarithmic slope of the IMF in the region around 1\\,\\msun, significantly flatter than the present-day value in the Milky Way disk of $x= 1.3\\pm 0.3$. The best-fitting luminosity-weighted formation redshift of the stars in massive cluster galaxies is $3.7^{+2.3}_{-0.8}$, and a possible interpretation is that the characteristic mass $m_c$ had a value of $\\sim 2$\\,\\msun\\ at $z\\sim 4$ (compared to $m_c \\sim 0.1$\\,\\msun\\ today), in qualitative agreement with models in which the characteristic mass is a function of the Jeans mass in molecular clouds. Such a ``bottom-light'' IMF for massive cluster galaxies has significant implications for the interpretation of measurements of galaxy formation and evolution. Applying a simple form of IMF evolution to literature data, we find that the volume-averaged star formation rate at high redshift may have been overestimated (by a factor of $3-4$ at $z> 4$), and the cosmic star formation history may have a fairly well-defined peak at $z\\sim 1.5$. The $M/L_V$ ratios of galaxies are less affected than their star formation rates, and future data on the stellar mass density at $z>3$ will provide further constraints on IMF evolution. The formal errors likely underestimate the uncertainties, and confirmation of these results requires a larger sample of clusters and the inclusion of redder rest-frame colors in the analysis. ", "introduction": "The form of the stellar initial mass function (IMF) is of fundamental importance for many areas of astrophysics and a topic of considerable debate (see, e.g., {Schmidt} 1959; {Miller} \\& {Scalo} 1979; {Scalo} 1986; {Larson} 1998, 2003; {Kroupa} 2002; {Chabrier} 2003, for reviews). Measurements of the IMF are difficult and somewhat model-dependent as they require the conversion of the observed present-day luminosity function of a stellar population to its mass function at birth. Best estimates for the Galactic disk suggest that the IMF has a powerlaw slope at $m\\gtrsim 1$\\,\\msun, and turns over at lower masses ({Kroupa} 2001; {Chabrier} 2003). This turnover can be modeled by a broken powerlaw ({Kroupa} 2001) or by a log-normal distribution with a characteristic mass $m_c$ ({Chabrier} 2003). The value of $m_c$ is $\\sim 0.1$\\,\\msun\\ in the disk of the Milky Way, with considerable uncertainty. The powerlaw slope at high masses is probably close to the {Salpeter} (1955) value of $x=1.35$, with an uncertainty of $\\sim 0.3$ ({Scalo} 1986; {Chabrier} 2003). Although there is no direct evidence for dramatic variations of the IMF within the present-day Milky Way disk (e.g., {Kroupa} 2001; {Chabrier} 2003), this does not preclude variations with time, metallicity, and/or environment. In particular, {Larson} (1998, 2005) has argued that the characteristic turnover mass may be largely determined by the thermal Jeans mass, which strongly depends on temperature ($\\propto T^{3/2}$ at fixed density). In the context of this model one might expect that heating by ambient far-infrared radiation would disfavor the formation of low mass stars in extreme environments, such as in super star clusters and in the center of the Milky Way. Other models emphasize the role of turbulence as opposed to temperature in determining the distribution of protostellar clumps (e.g., {Padoan} \\& {Nordlund} 2002), and in these models the role of the environment may be less direct (see {McKee} \\& {Ostriker} [2007] for a recent review of various models to explain the characteristics of the IMF). Observations may support the notion of a top-heavy (or ``bottom-light'') IMF in extreme environments. Some young super star clusters in M82 appear to have a top-heavy mass function (e.g., {Rieke} {et~al.} 1993; {McCrady}, {Gilbert}, \\& {Graham} 2003), as do clusters in the Galactic center region (e.g., {Figer} {et~al.} 1999; {Stolte} {et~al.} 2005; {Maness} {et~al.} 2007). The interpretation of observed mass functions is complicated by dynamical effects, which tend to make the mass function more top-heavy over time, in particular in the central regions of star clusters (see, e.g., {McCrady} {et~al.} 2003; {McCrady}, {Graham}, \\& {Vacca} 2005; {Kim} {et~al.} 2006). Recently Harayama, Eisenhauer, \\& Martins (2007) studied the IMF of NGC 3603, one of the most massive Galactic star-forming regions, out to large radii and conclude that its IMF is substantially flatter than Salpeter for masses $0.4 -20$\\,\\msun. The IMF may also depend on redshift. At earlier times star formation presumably occurred more often in a burst mode than in a relatively gradual ``disk'' mode (e.g., {Steidel} {et~al.} 1996; {Blain} {et~al.} 1999b; {Lacey} {et~al.} 2007), which means that the IMF could generally be more skewed toward high mass stars at redshifts 1--3 and beyond. Furthermore, the average metallicity in star forming clouds was lower at higher redshift, which may have led to an extremely top-heavy IMF for the first generation of stars (e.g., {Abel}, {Bryan}, \\& {Norman} 2002; {Bromm}, {Coppi}, \\& {Larson} 2002). Finally, the cosmic microwave background (CMB) radiation sets a floor to the ambient temperature, and hence the Jeans mass, which scales with $(1+z)$. Beyond $z\\sim 2$ the CMB temperature exceeds the typical temperatures of dense prestellar cores in Galactic molecular clouds (e.g. {Evans} {et~al.} 2001; {Tafalla} {et~al.} 2004). Therefore, at sufficiently high redshift the characteristic mass may be expected to evolve roughly as $m_c \\propto (1+z)^{3/2}$, leading to IMFs which have a reduced fraction of low mass stars ({Larson} 1998). The effects of the CMB are even more pronounced when its influence on the pressure in star-forming clouds is taken into account, and {Larson} (2005) suggests that at $z=5$ the characteristic mass may be higher than today's value by as much as an order of magnitude. Such rapid evolution of the IMF would have important consequences for determinations of masses and star formation rates of distant galaxies, and for measurements of evolution in these properties. It is very difficult to constrain the IMF at early times directly, as the light of high redshift galaxies is completely dominated by massive stars. The extremely blue rest-frame UV colors of galaxies at $z\\sim 6$ may imply a top-heavy IMF ({Stanway}, {McMahon}, \\& {Bunker} 2005), although this is just one of several possible explanations. From observations of a lensed Lyman break galaxy there is some evidence that the {\\em slope} of the IMF at $z\\sim 3$ is similar to the Salpeter value at the high mass end ({Pettini} {et~al.} 2000), but there is essentially no information on stars with masses near or below 1\\,\\msun. Fortunately, the form of the high redshift IMF has implications for the properties of galaxies at much lower redshift, as all stars with masses $\\lesssim 0.8$\\,\\msun\\ that formed in the history of the Universe are still with us today. Tumlinson (2007) finds that the properties of carbon-enhanced metal-poor stars in our Galaxy are best explained with a relatively high number of stars in the mass range 1--8\\,\\msun\\ at high redshift. Various other constraints obtained from galaxies at low redshift (including our own) are reviewed in {Chabrier} (2003). Of particular interest are the stellar populations of massive early-type galaxies, as they are very homogeneous and should reflect conditions in star forming regions at $z> 2$. Recently, {Cappellari} {et~al.} (2006) used the kinematics of elliptical galaxies to constrain the IMF, as the dynamical $M/L$ ratio provides an upper limit to the amount of mass that can be locked up in low mass stars. Current data appear to rule out a {Salpeter} (1955) (or steeper) IMF, but are consistent with {Kroupa} (2001) and {Chabrier} (2003) IMFs ({Cappellari} {et~al.} 2006). In this paper, we provide new constraints on the IMF at high redshift by comparing the {\\em evolution} of the $M/L$ ratios of early-type galaxies to their color evolution. This method was first suggested by {Tinsley} (1980), but data of sufficient accuracy are only now becoming available. The method is sensitive to the IMF in the important mass range around 1\\,\\msun, where the effects of an evolving characteristic mass might be expected to manifest themselves. A plan of the paper follows. In \\S\\,2 a relation between color evolution, luminosity evolution, and the logarithmic slope of the IMF $x$ is derived using stellar population synthesis models. In \\S\\,3 published data and archival HST images of galaxy clusters are used to construct the redshift evolution of the $U-V$ color-mass relation. In \\S\\,4 the color evolution from \\S\\,3 is combined with the previously measured evolution of the $M/L_B$ ratio. The relations from \\S\\,2 are then used to derive constraints on the IMF slope $x$ from the combined color and luminosity evolution. Section 5 is devoted to the (many) systematic uncertainties in the methodology and in the data, and \\S\\,6 asks whether our results are consistent with other constraints on the stellar populations of massive early-type galaxies. Although the constraints we derive in this paper are subject to many uncertainties, it is interesting to explore their consequences. In \\S\\,7 the fitting results of \\S\\,4 are interpreted in the context of an evolving characteristic mass $m_c$. The data on cluster galaxies are combined with previous constraints on $m_c$ for globular clusters and submm-galaxies, and a simple form of IMF evolution is proposed. This evolution is then applied to literature data on the evolution of the volume-averaged star formation rate and stellar mass density. The key results are summarized in \\S\\,8. We assume $\\Omega_m=0.3$, $\\Omega_{\\Lambda}=0.7$, and $H_0=71$\\,\\kms\\,Mpc$^{-1}$ where needed. ", "conclusions": "\\label{conclusion.sec} This paper compares the color evolution of massive cluster galaxies to their luminosity evolution, with the aim of constraining the form of the IMF at the time when the stars in these galaxies were formed. It is found that the evolution of the rest-frame $U-V$ color is not consistent with the previously determined evolution of the rest-frame $M/L_B$ ratio, unless the IMF slope is significantly flatter than the Salpeter value around 1\\,\\msun. For standard IMFs with a slope of 1.3 at $m\\geq 1$\\,\\msun\\ the luminosity evolution is too fast for the measured color evolution, and the implied stellar ages derived from $M/L$ evolution and color evolution are not consistent with each other. The only models that are able to fit the color evolution and the luminosity evolution simultaneously have IMF slopes of $\\sim 0$ around 1\\,\\msun\\ and mean luminosity-weighted stellar formation redshifts of $\\sim 4$ (for Solar metallicity). This result is somewhat uncertain, as the currently available sample of cluster galaxies with accurate rest-frame $U-V$ colors and dynamical masses is somewhat limited and there are many systematic effects which may play a role. In particular, it is an open question whether stellar population synthesis models are able to predict color evolution with the required accuracy. The commonly used {Bruzual} \\& {Charlot} (2003) and {Maraston} (2005) models give broadly similar answers, but that may be because they share many of the same assumptions. As discussed in \\S\\,\\ref{comp.sec} the higher stellar ages implied by a flat IMF are consistent with many other studies, which lends some credibility to the results presented here. Of particular importance is the agreement with the data on Balmer line strengths of {Kelson} {et~al.} (2001), as they do not suffer from the same systematic uncertainties as the color data. Formation redshifts substantially larger than two also fit more comfortably with the direct detection of old galaxies at high redshifts. A firm independent measurement of the star formation epoch of massive cluster galaxies, combined with a better understanding of selection effects at high redshift, would leave the IMF as the only free parameter and greatly simplify the problem. The implications discussed in \\S\\,\\ref{imply.sec} are obviously somewhat speculative. Although the interpretation in terms of an evolving characteristic mass is physically plausible according to some models (e.g., {Larson} 2005), many other forms of the IMF are consistent with the data. The observations described in this paper are only sensitive to a narrow mass range near 1\\,\\msun, and the IMF proposed in Eq.\\ \\ref{mod.eq} represents a very substantial extrapolation. This is illustrated in Fig.\\ \\ref{summary.plot}: both the top-heavy IMF (red line) and the ``bottom-light'' IMF (green line) are consistent with the data presented in this paper. The main reason for preferring the bottom-light IMF over a top-heavy form is that the absolute $M/L$ ratios of galaxies are within a reasonable range. As shown in Fig.\\ \\ref{mlmc.plot} the $M/L_V$ ratios are similar to those implied by a standard {Chabrier} (2003) IMF, which means that they are consistent with dynamical measurements at $z=0$ ({Cappellari} {et~al.} 2006). \\vbox{ \\begin{center} \\leavevmode \\hbox{% \\epsfxsize=8.5cm \\epsffile{f15.eps}} \\figcaption{\\small Illustration of the key results of this paper. In \\S\\,4 the slope of the IMF of massive cluster galaxies $x$ is estimated to be approximately $-0.3$ in a narrow region around 1\\,\\msun\\ (thick black line). The red dashed line and the green solid line show two possible interpretations: a global change of the slope of the IMF (with respect to a standard Chabrier or Salpeter IMF) at all masses (red dashed line), or a change in the characteristic mass (solid green line). The ``top-heavy'' interpretation is inconsistent with the dynamical $M/L$ ratios of nearby elliptical galaxies (see \\S\\,\\ref{absml.sec}), whereas the ``bottom-light'' interpretation is consistent with all data that we are aware of. The blue and yellow areas illustrate that stars with masses $\\gtrsim 10$\\,\\msun\\ drive star formation measurements, whereas stars with masses $1-5$\\,\\msun\\ drive $M/L$ measurements (see \\S\\,7.3-7.5). \\label{summary.plot}} \\end{center}} An ``unintended'' effect of an evolving IMF of the form proposed here is that it reduces the discrepancy between the observed stellar mass density and the density implied by the cosmic star formation history. This result is in excellent (albeit qualitative) agreement with several other recent studies (e.g., Hopkins \\& Beacom 2006, Fardal et al.\\ 2007, P\\'erez-Gonz\\'alez et al.\\ 2007; Wilkins, Trentham, \\& Hopkins 2007; Dav\\'e 2007; see also, e.g., Fields 1999). The differences between a non-evolving IMF and an evolving IMF are fairly large at $z\\sim 4$ (as the effect on the star formation rate is strong and the effect on $M/L$ ratios is weak at that redshift), and it will be interesting to see where future measurements of the mass density at high redshift will fall in Fig.\\ \\ref{mdens.plot}. We note that the effects on the $M/L_V$ ratios are somewhat uncertain, as they rely on rather rudimentary stellar population synthesis modeling. The effects on star formation rates are more robust, and suggest that the cosmic star formation rate peaked at $z\\sim 1.5$. This paper adds to previous theoretically and observationally motivated suggestions that the IMF may evolve with redshift (e.g., Worthey et al.\\ 1992; Larson 1998, 2005; Fields 1999; Blain et al.\\ 1999a; Baugh et al.\\ 2005; Stanway et al.\\ 2005; Hopkins \\& Beacom 2006; Tumlinson 2007; Lacey et al.\\ 2007; Fardal et al.\\ 2007; P\\'erez-Gonz\\'alez et al.\\ 2007). Although these studies vary greatly in their parameterization of IMF evolution and the range of stellar masses that are considered, they all suggest that the ratio of high-mass stars to low-mass stars was higher in the past. It should be pointed out that most of these papers invoke a change in the IMF as a ``last resort'' possibility, to explain data that are otherwise difficult to interpret. In the present study a different approach was followed, in that we set out with the specific purpose of constraining the slope of the IMF. An advantage of the applied method is that it is fairly direct, as the rate of luminosity evolution is determined by the number of stars as a function of mass. Disadvantages are that it is only sensitive to a very limited mass range (see Fig.\\ \\ref{summary.plot}); that it relies on stellar population synthesis models, which are not well calibrated in the relevant parameter range; that the progenitors of early-type galaxies may not be representative for the general population of high redshift galaxies; and that the currently available data are somewhat limited. Accepting the possibility of an evolving characteristic mass, it is interesting to speculate what could be the cause, or causes. The proximate cause may well be a higher temperature in molecular clouds at high redshift, which would raise the Jeans mass and could inhibit the formation of low mass stars ({Larson} 1998, 2005). The ultimate cause could be the higher temperature of the cosmic microwave background, the fact that star formation tends to proceed in more extreme environments at higher redshift, or a combination. Available information on IR-bright galaxies suggests that dust temperatures in star burst galaxies are of order $30-40$\\,K ({Dunne} {et~al.} 2000; {Chapman} {et~al.} 2005), and hence exceed the CMB temperature for all relevant redshifts. However, it is as yet unclear what fraction of the total star formation has taken place in these extreme environments (see, e.g., {Reddy} {et~al.} 2007). The analysis presented here can be improved in various ways. The number of clusters with accurate rest-frame $U-V$ is currently smaller than the number of clusters with accurate $M/L_B$ measurements, and this can be remedied by obtaining accurate (space-based) photometry in well-chosen filters of the remaining clusters in the vv07 sample. It is also important to measure the evolution in a redder rest-frame color, such as $V-I$. Redder color suffer less from possible contributions of hot stars, and their evolution is probably somewhat better calibrated in stellar population synthesis models. This requires very accurate photometry in the near-infrared, which should be possible with WFC3 on HST. On the modeling side, it would be helpful to implement more variations of the IMF in stellar population synthesis codes than the standard Salpeter, Kroupa, and Chabrier forms. Ultimately it may be fruitful (or prove necessary) to have the characteristic mass, or some other parameter describing the form of the IMF, as one of the ``standard'' parameters in these models, on a par with the age and metallicity." }, "0710/0710.2041_arXiv.txt": { "abstract": "\\noindent The generalized Darmois--Israel formalism for Einstein--Gauss--Bonnet theory is applied to construct thin-shell Lorentzian wormholes with spherical symmetry. We calculate the energy localized on the shell, and we find that for certain values of the parameters wormholes could be supported by matter not violating the energy conditions. ", "introduction": "For spacetime dimension $D\\geq 5$ the Einstein--Hilbert action of gravity admits quadratic corrections constructed from coordinate-invariant tensors which scale as fourth derivatives of the metric. In particular, when $D=5$ the most general theory leading to second order equations for the metric is the so-called Einstein--Gauss--Bonnet theory or Lovelock theory up to second order. This class of model for higher dimensional gravity has been widely studied, in particular because it can be obtained in the low energy limit of string theory \\cite{1}. For spacetime dimensions $D<5$ the Gauss--Bonnet terms in the action represent a topological invariant. The equations of gravitation admit solutions, known as Lorentzian wormholes, which connect two regions of the same universe (or of two universes) by a throat, which is a minimal area surface \\cite{motho,book}. Such kind of geometries would present some features of particular interest, as for example the possibility of time travel (see Refs. \\cite{morris-novikov}). But a central objection against the actual existence of wormholes is that in Einstein gravity the flare-out condition \\cite{hovis1} to be satisfied at the throat requires the presence of exotic matter, that is, matter violating the energy conditions \\cite{book}. In this sense, thin-shell wormholes have the advantage that the exotic matter would be located only at the shell. However, it has recently been shown \\cite{gra-wi} that in pure Gauss--Bonnet gravity exotic matter is no needed for wormholes to exist; in fact, they could exist even with no matter (see also Refs. \\cite{7,maeda}). In this work we thus study thin-shell wormholes in Einstein--Gauss--Bonnet gravity. We focus in the amount of matter necessary for supporting the wormholes, without analyzing the microphysics explaining this matter. Differing from the approach in the related work Ref. \\cite{grg06}, where the Gauss--Bonnet terms were treated as an effective source for the Einstein's field equations, here the Gauss--Bonnet contribution is treated as an essentially geometrical object. This requires a generalization \\cite{4} of the Darmois--Israel formalism \\cite{3} for thin shells, but provides a better physical understanding. In particular, we show that for certain values of the parameters, thin-shell wormholes could be supported by matter not violating the energy conditions. ", "conclusions": "Motivated by the results within pure Gauss--Bonnet gravity (i.e. without Einstein term) in Ref. \\cite{gra-wi}, here we evaluate the amount of exotic matter and the energy conditions, following the approach presented above in which the Gauss--Bonnet term is treated as a geometrical contribution in the field equations. Coming this contribution from the curvature tensor, this approach is clearly the most suitable to give a precise meaning to the characterization of matter supporting the wormhole. As we shall see, the results will considerably differ from those in Ref. \\cite{grg06}, where the Gauss--Bonnet term was treated as an effective source for the Einstein's field equations. The {\\it weak energy condition} (WEC) states that for any timelike vector $U^{\\mu}$ it must be $T_{\\mu\\nu}U^{\\mu}U^{\\nu}\\geq0$; the WEC also implies, by continuity, the {\\it null energy condition} (NEC), which means that for any null vector $k^{\\mu}$ it mus be $T_{\\mu\\nu}k^{\\mu}k^{\\nu}\\geq0$ \\cite{book}. In an orthonormal basis the WEC reads $\\rho\\geq0$, $\\rho+p_{l}\\geq0\\ \\forall\\, l$, while the NEC takes the form $\\rho+p_{l}\\geq0 \\ \\forall\\, l$. In the case of thin-shell wormholes the radial pressure $p_{r}$ is zero, while within Einstein gravity or even with the inclusion of a Gauss--Bonnet term in the way proposed in \\cite{grg06}, the surface energy density must fulfill $\\sigma<0$, so that both energy conditions would be violated. The sign of $\\sigma+p_{t}$ where $p_{t}$ is the transverse pressure is not fixed, but it depends on the values of the parameters of the system. In what follows we restrict to static configurations. The surface energy density $\\sigma_{0}$ and the transverse pressure $p_{0}$ for a static configuration ($b=b_0$, $\\dot{b}=0$, $\\ddot{b}=0$) are given by \\begin{equation} \\sigma_{0}=-\\frac{1}{8\\pi}\\left[ 6\\frac{\\sqrt{f(b_{0})}}{b_{0}}-2\\alpha\\sqrt{f(b_{0})}\\left(4\\frac{f(b_{0})}{b^{3}_{0}}-\\frac{12}{b^{3}_{0}}\\right)\\right],\\label{s0} \\end{equation} \\begin{equation} p_{0}=\\frac{1}{8\\pi}\\left[ 4\\frac{\\sqrt{f(b_{0})}}{b_{0}}+\\frac{f^{'}(b_{0})}{\\sqrt{f(b_{0})}}-2\\alpha\\left(2f^{'}(b_{0})\\frac{\\sqrt{f(b_{0})}}{b^{2}_{0}} -2\\frac{f^{'}(b_{0})}{b^{2}_{0}\\sqrt{f(b_{0})} }\\right)\\right].\\label{p0} \\end{equation} Note that the sign of the surface energy density is, in principle, not fixed. The most usual choice for quantifying the amount of exotic matter in a Lorentzian wormhole is the integral \\cite{nandi}: \\begin{equation} \\Omega= \\int (\\rho + p_{r})\\sqrt{|g|}\\,d^{4}x. \\end{equation} We can introduce a new radial coordinate ${\\cal R}=\\pm(r-b_{0})$ with $\\pm$ corresponding to each side of the shell. Then, because in our construction the energy density is located on the surface, we can write $\\rho=\\delta({\\cal R})\\,\\sigma_{0}$, and because the shell does not exert radial pressure the amount of exotic matter reads \\begin{equation} \\Omega=\\int^{2\\pi}_{0} \\int^{\\pi}_{0}\\int^{\\pi}_{0}\\int^{+\\infty}_{-\\infty}\\delta({\\cal R})\\,\\sigma_{0} \\sqrt{|g|}\\, d{\\cal R}\\,d\\theta\\,d\\chi\\ d\\varphi =2\\pi^{2} b_{0}^3 \\sigma_{0}. \\end{equation} Replacing the explicit form of $\\sigma_{0}$ and $g$, we obtain the exotic matter amount as a function of the parameters that characterize the configurations: \\begin{equation} {\\Omega}= -\\frac{3}{2}\\pi b^{2}_{0}\\sqrt{f(b_{0})} +2\\pi\\alpha\\sqrt{f(b_{0})}\\left[f(b_{0})-3)\\right], \\end{equation} where $f$ is given by (\\ref{f}). For $\\Lambda=0$, in the limit $\\alpha\\rightarrow 0$ and Taylor expanding up to zeroth order we obtain the exotic matter for the Reissner--N\\\"{o}rdstrom ($Q\\neq 0$) and Schwarzschild ($Q= 0$) geometries. For $\\alpha\\neq 0$ we now find that there exist positive contributions to $\\Omega$; these come from the different signs in the expression (\\ref{s0}) for the surface energy density, because $\\Omega$ is proportional to $\\sigma_0$. We stress that this would not be possible if the standard Darmois--Israel formalism was applied, treating the Gauss--Bonnet contribution as an effective energy-momentum tensor, because this leads to $\\sigma_0\\sim - \\sqrt{f(b_0)}/b_0$ \\cite{grg06}. Now, once the explicit form of the function $f$ (with $\\Lambda=0$) is introduced, the condition $\\sigma_0>0$ leads to \\begin{equation} -8\\alpha-2b_0^2-b_0^2\\sqrt{1+\\frac{16M\\alpha}{\\pi b_0^4}-\\frac{8Q^2\\alpha}{3b_0^6}}>0,\\label{s+} \\end{equation} which can hold only for $\\alpha <0$. The subsequent analysis is simplified by considering the case $Q=Q_c, \\,\\Lambda=0$; then there would be at most only one horizon in the original manifold, its radius being independent of the charge. But for this charge it can be shown that, for values of $\\alpha$ such that the horizon exists, it is not possible to fulfil Eq. (\\ref{s+}) for any wormhole radius larger than $r_{hor}$; the reason is that the horizon exists only for $\\alpha > -M/(3\\pi)$, which is not compatible with condition (\\ref{s+}) if $b_0$ is to be larger than the corresponding horizon radius. Instead, a simple numerical analysis shows that for $\\alpha$ slightly below $-M/(3\\pi)$ both the singularity at $r\\neq 0$ and the horizon dissapear in the original manifold, so that the only condition to be fulfilled is that given by Eq. (\\ref{s+}). And for $\\alpha <0$ it is always possible to choose $b_0$ such that this indeed happens, so that the existence of thin-shell wormholes is compatible with a positive surface energy density \\footnote{It is not difficult to see that for $Q$ slightly below $Q_c$ the same happens for larger $|\\alpha|$.}. In figures 1 and 2 we show the amount $\\Omega$ as a function of the wormhole radius for this relatively large value of $|\\alpha|$ (that is, $\\pi |\\alpha|$ of order $M$); though this would imply microscopic configurations or a scenario different from that suggested by present day observation, the analysis shows that this is the most interesting situation. Besides the fact that $\\Omega$ results to be smaller when calculated by treating the Gauss--Bonnet contribution as a geometric object than in the case that it was treated as an effective energy-momentum tensor, this amount is smaller than which would be necessary in the five-dimensional pure Reissner--N\\\"{o}rdstrom case (see Fig. 1). However, the central, remarkable, result is that we have a region with $\\Omega >0$ (see Fig. 2), corresponding to $\\sigma_0>0$; and that besides, from Eqs. (\\ref{s0}) and (\\ref{p0}) we have $\\sigma_0+p_0=-(b_0/3)\\,d\\sigma_0/db_0$, which shows that for wormhole radii such that $\\sigma_0 >0$ and ${d\\sigma_0}/{db_0}<0$ ($r_{wh}$ within the maximun and the zero of $\\sigma_0$ in Fig. 3) both the WEC and the NEC are satisfied. Thus, in the picture providing a clear meaning to matter in the shell, in Einstein--Gauss--Bonnet gravity the violation of the energy conditions could be avoided, and wormholes could be supported by ordinary matter. \\begin{figure}[htp] \\centering \\includegraphics[height=8cm, width=12cm]{compara.eps} \\caption{The amount $\\Omega$ is shown as a function of $r^2_{wh}/M$, for $Q=Q_c$ and $\\alpha=-0.11 M$. The dashed line corresponds to the five-dimensional Reissner--N\\\"{o}rdstrom case, the dotted line corresponds to considering the Gauss--Bonnet term as a kind of effective source for the field equations, and the solid line shows the result obtained here with the generalized Darmois--Israel formalism for Einstein--Gauss--Bonnet theory.} \\end{figure} \\begin{figure}[htp] \\centering \\includegraphics[height=8.cm, width=8.5cm]{omega.eps} \\caption{The amount $\\Omega$ is shown as a function of $r^2_{wh}/M$, for $Q=Q_c$ and $\\alpha=-0.11 M$. The plot shows the result obtained here with the generalized Darmois--Israel formalism for Einstein--Gauss--Bonnet theory.} \\end{figure} \\begin{figure}[htp] \\centering \\includegraphics[height=8cm, width=8cm]{wec.eps} \\caption{Energy conditions: the dashed line shows $\\tilde \\sigma_0=\\sqrt{M}\\sigma_0$, the dashed-dotted line shows $\\tilde p_0=\\sqrt{M}p_0$ and the solid line shows the sum $\\tilde \\sigma_0+\\tilde p_0$.} \\end{figure}" }, "0710/0710.1873_arXiv.txt": { "abstract": "Precision measurement of the scalar perturbation spectral index, $n_s$, from the \\emph{Wilkinson Microwave Anisotropy Probe} temperature angular power spectrum requires the subtraction of unresolved point source power. Here we reconsider this issue, attempting to resolve inconsistencies found in the literature. First, we note a peculiarity in the \\emph{WMAP} temperature likelihood's response to the source correction: Cosmological parameters do not respond to increased source errors. An alternative and more direct method for treating this error term acts more sensibly, and also shifts $n_s$ by $\\sim0.3\\sigma$ closer to unity. Second, we re-examine the source fit used to correct the power spectrum. This fit depends strongly on the galactic cut and the weighting of the map, indicating that either the source population or masking procedure is not isotropic. Jackknife tests appear inconsistent, causing us to assign large uncertainties to account for possible systematics. Third, we note that the \\emph{WMAP} team's spectrum was computed with two different weighting schemes: uniform weights transition to inverse noise variance weights at $l=500$. The fit depends on such weighting schemes, so different corrections apply to each multipole range. For the Kp2 mask used in cosmological analysis, we prefer source corrections {$A=0.012\\pm0.005$ $\\mu$K$^2$} for uniform weighting and {$A=0.015\\pm0.005$ $\\mu$K$^2$} for $N_{\\rm obs}$ weighting. Correcting \\emph{WMAP}'s spectrum correspondingly, we compute cosmological parameters with our alternative likelihood, finding $n_s=0.970\\pm0.017$ and $\\sigma_8=0.778\\pm0.045$. This $n_s$ is only $1.8\\sigma$ from unity, compared to the $\\sim 2.6\\sigma$ \\emph{WMAP} 3-year result. Finally, an anomalous feature in the source spectrum at $l<200$ remains, most strongly associated with W-band. ", "introduction": "Measuring $n_s$, the spectral index of initial scalar fluctuations, which is scale invariant ($n_s = 1$) in the Harrison-Zeldovich model and slightly shallower in inflation models, is difficult, primarily because experimental systematics require control over a broad range of spatial scales. In inflation, the deviation from unity closely relates to the inflationary potential and the number of $e$-folds of expansion, so a statistically robust measurement of $n_s \\neq 1$ places compelling constraints on the physics of the inflationary epoch. Because all-sky measurements of the cosmic microwave background (CMB) access the largest observable scales in the universe, the angular power spectrum of the CMB, with a long lever arm, is crucial to such studies. Indeed, the latest data release from the \\textit{Wilkinson Microwave Anisotropy Probe} (\\emph{WMAP}) claims $\\sim 2.6\\sigma$ deviation from the Harrison-Zeldovich spectrum \\citep{Spergel2007}. Unfortunately, the CMB is not a totally clean measurement. For example, the well-known degeneracy with the optical depth since recombination ($\\tau$) makes precision measurement of $n_s$ impossible using CMB temperature anisotropies alone, and polarization is required to break it. Complicated noise properties and hints of unknown systematics in the \\emph{WMAP} measurement of large-scale polarization indicate that the systematic uncertainty in both $\\tau$ and $n_s$ should still be considered significant \\citep{Eriksen2007b}. Another important, but under-appreciated, complication for the measurement of $n_s$ is additional power in the angular spectrum from unresolved, and unmasked, point sources. At high $l$, this shot noise can significantly bias the power spectrum, and consequently $n_s$. The \\emph{WMAP} team devised a sensible prescription for dealing with this contaminant: 1) Use the spectral energy distribution measured from detected sources (and distinct from the CMB) to infer it for undetected ones; 2) measure the contamination using multifrequency data; 3) correct the spectrum; and 4) marginalize over the measurement error when computing the likelihood \\citep{Hinshaw2003,Hinshaw2007}. \\citet{Huffenberger2004} found a level of source contamination consistent with the level in the first \\emph{WMAP} data release \\citep{Hinshaw2003}. However, based on the three year temperature data \\citep{Hinshaw2007}, \\citet{Huffenberger2006} measured a point source spectrum with two irregularities. First, at $l>200$ the spectrum is white, but with an amplitude below the value in the original preprint of \\citet{Hinshaw2007}. In the present work, we discovered a small error in the power spectra used for the \\citet{Huffenberger2006} estimate, which should have reported $A = 0.013 \\pm 0.001$ $\\mu$K$^2$ instead of $A = 0.011 \\pm 0.001$ $\\mu$K$^2$, still below the original WMAP value of $A = 0.017 \\pm 0.002$ $\\mu$K$^2$. Prompted by our result, \\citet{Hinshaw2007} re-examined the issue, revising their value down somewhat and increasing the error bars, to $A = 0.014 \\pm 0.003$ $\\mu$K$^2$. The \\citet{Spergel2007} bispectrum analysis indicates a non-Gaussianity consistent with these values, but lacks the statistical power of the multifrequency power spectrum comparison. The second peculiarity is that the power at $100 80$ \\AA, we eliminate all but the strongest (and rarest) [O~II] emitters \\citep{gronw07}. The narrow-band imaging in ECDF-S is supplemented by optical and near-IR observations from MUSYC \\citep{gawis06a}, GOODS \\citep{dicki03}, and GEMS \\citep{rix04} surveys, and Spitzer observations from GOODS and SIMPLE\\footnotemark\\ (Damen et al., in preparation). Using the multi-wavelength data available in ECDF-S, we study the rest-frame UV to near-IR properties of this large sample of LAEs. We place special emphasis on the new Spitzer IRAC observations, which sample the rest-frame $0.9 - 2$ \\micron\\ emission from the $z = 3.1$ LAEs and provide valuable constraints on their stellar population and star formation history. \\footnotetext{http://www.astro.yale.edu/dokkum/SIMPLE/} We assume a ${\\rm \\Lambda}$CDM cosmology with $\\Omega_{\\rm M} = 0.3$, $\\Omega_{\\rm \\Lambda} = 0.7$, and $h = 0.7$. All magnitudes are in the AB system. ", "conclusions": "\\label{Disc} There are several previous studies on the stellar population of LAEs at $z \\sim 3$ and beyond \\citep{gawis06b, finke07, pirzk07, nilss07}. These previous studies found that LAEs have masses ranging from $10^6 - 10^9$ \\Msun, and ages ranging from $0.005 - 1$ Gyr. Our best-fit average age of 160 Myr and mass of \\pow{3}{8} \\Msun\\ for the IRAC-undetected sample are broadly consistent with these previous studies. In a further analysis of the IRAC-undetected sample presented in this paper, \\citet[submitted]{gawis07} fit two-component models and find that the IRAC-undetected LAEs have a total mass of \\pow{1}{9} \\Msun, with a young stellar component (age $\\sim 20$ Myr) accounting for $\\sim 20\\%$ of the total mass. These results are consistent with ours which are derived using single-component models and should be interpreted as the average properties of the entire stellar population within the LAEs. \\begin{figure} \\includegraphics[width=\\columnwidth]{f4.eps} \\caption{ A comparison of the $R-m_{3.6}$ colors and 3.6 \\micron\\ magnitudes between 8 \\micron\\ detected LBGs (green square), regular LBGs (black dots), and the IRAC-detected LAEs (red stars). The solid line shows the $R<25.5$ selection limit for LBGs. The data for the regular LBGs come from IRAC observations of the \\citet{steid03} LBG sample (Magdis et al., in preparation), and the 8 \\micron\\ LBG data come from combining the samples of Magdis et al.\\ and \\citet{rigop06}. The mean colors and magnitudes of each sample are plotted as large filled symbols, with error bars showing the dispersion within the samples. We additionally plot the stacked photometry of the IRAC-undetected sample, shown as the red filled red star with the faintest 3.6 \\micron\\ magnitude. \\label{compLBG} } \\end{figure} The IRAC-detected LAEs are more luminous than the IRAC-undetected LAEs in both rest-frame UV and near-IR, and represent the massive end of the LAE mass spectrum. Potentially, the IRAC-detected LAEs may provide a link between the LAEs and other galaxy populations. For instance, they may be the $z \\sim 3$ analogues to the similarly massive IRAC-detected LAEs found at $z \\sim 5.7$ \\citep{lai07}. In \\Fig{compLBG}, we compare the $R-m_{3.6}$ colors and 3.6 \\micron\\ magnitudes of the LAEs to 8 \\micron\\ detected LBGs and regular LBGs at $z \\sim 3$. The $R-m_{3.6}$ color and 3.6 \\micron\\ magnitude correlate with stellar mass \\citep{rigop06}, with the 8 \\micron\\ LBGs being the most massive ($\\ga 10^{11}$ \\Msun), followed by LBGs ($\\sim 10^{10} - 10^{11}$ \\Msun), IRAC-detected LAEs ($\\sim 10^{10}$ \\Msun), and finally IRAC-undetected LAEs ($\\sim 10^8$ \\Msun). The IRAC-detected LAEs occupy the faint and blue end of the LBG color-magnitude distribution, suggesting that they may be the lower mass extension of the LBG population. This interpretation is supported by the fact that the inferred mass of the IRAC-detected LAEs ($\\sim 10^{10}$ \\Msun) is at the low end of the mass range found for LBGs ($\\sim 10^{10} - 10^{11}$ \\Msun; \\citealt{papov01, shapl01}). Furthermore, most LAEs in the present sample have (observer frame) optical colors that would allow them to be selected as LBGs, although the LAEs tend to be fainter in the rest-frame UV \\citep{gronw07, gawis06b}. It has also been observed that $\\sim 20\\% - 25\\%$ of LBGs at $z \\sim 3$ exhibit Ly$\\alpha$ emission strong enough to be narrow band excess objects \\citep{steid00, shapl03}. There is thus accumulating evidence suggesting that the IRAC-detected LAEs may be the bridge connecting the LAE and LBG populations. Using the stellar mass estimates derived in the previous section and assuming a survey volume of \\pow{1.1}{5} Mpc$^3$ \\citep{gronw07}, we find stellar mass densities of $0.3 \\pm 0.3 \\times 10^6$ and $3 \\pm 1 \\times 10^6$ \\Msun\\ Mpc$^{-3}$ for the IRAC-undetected and IRAC-detected populations, respectively. The errors in the stellar mass densities include uncertainties in the mass from the fits and uncertainties in the number density coming from Poisson fluctuations and a $\\sim 20\\%$ cosmic variance given the observed LAE bias of $\\sim 1.7$ \\citep[submitted]{gawis07}. We should stress that the above stellar mass densities only account for LAEs with Ly$\\alpha$ luminosities above our survey completeness limit of \\pow{1.3}{42} erg s$^{-1}$. The IRAC-undetected LAEs account for about $9 \\pm 10\\%$ of the total stellar mass in LAEs, even though they make up 2/3 of the total by number. Compared to LBGs and DRGs at $z \\sim 3$, which have stellar mass densities of $\\sim \\pow{1}{7}$ and \\pow{7}{6} \\Msun\\ Mpc$^{-3}$ respectively \\citep{grazi07}, the stellar mass contained in LAEs is smaller, but not insignificant. However, it is important to keep in mind that there may be substantial overlap between the LAEs and LBGs. A better understanding of the overlap between these two populations is necessary before a direct comparison of the stellar mass densities can be made. The present results show that the LAEs posses a wide range of masses and ages, from the massive and evolved IRAC-detected LAEs to the young and small IRAC-undetected LAEs. The range of photometric properties shown in the LAE sample suggests that LAEs exhibit a continuum of properties between these two extremes. The presence of both young and evolved stellar populations within the overall LAE population implies that the Ly$\\alpha$ luminous phase of galaxies may last $\\ga 1$ Gyr, or that the Ly$\\alpha$ luminous phase is recurring. Interestingly, \\citet{shapl01} found that LBGs with best-fit stellar population ages $\\ga 1$ Gyr also show strong Ly$\\alpha$ emission in their spectra. The authors suggest that vigorous past star formation has destroyed and/or expelled the dust inside the galaxies, allowing the Ly$\\alpha$ photons to escape. One characteristic of these evolved Ly$\\alpha$ emitting LBGs is that they tend to have more quiescent SFR than the rest of the population. The IRAC-detected LAEs, with their evolved stellar population and low specific SFR, are consistent with this scenario. If the LAE phase is recurring, then the young age of the IRAC-undetected LAEs implies that their number and stellar mass densities can be as much as a factor $\\sim 10$ higher (very roughly the ratio of the age of the universe to the stellar age). Similarly young stellar populations ($\\la 100$ Myr) have also been found at higher redshifts \\citep{finke07, pirzk07, verma07}. The stochastic nature of galaxies with short lifetimes suggests that there may be a related population of undetected pre or post-starburst galaxies that may contribute significantly to the stellar mass and star formation rate densities at $z \\ga 3$." }, "0710/0710.1730_arXiv.txt": { "abstract": "The evolution of stellar collision products in cluster simulations has usually been modelled using simplified prescriptions. Such prescriptions either replace the collision product with an (evolved) main sequence star, or assume that the collision product was completely mixed during the collision. It is known from hydrodynamical simulations of stellar collisions that collision products are not completely mixed, however. We have calculated the evolution of stellar collision products and find that they are brighter than normal main sequence stars of the same mass, but not as blue as models that assume that the collision product was fully mixed during the collision. ", "introduction": "The aim of the MODEST collaboration \\citep{article:modest1} is to model and understand dense stellar systems, which requires a good understanding of what happens when two single stars or binary systems undergo a close encounter. A possible outcome of such an encounter is a collision followed by the merging of two or more stars. This is a possible formation channel for blue straggler stars (\\emph{e.g.} \\citet{article:sills_on_axis}). Understanding the formation and evolution of blue stragglers is important for understanding the Hertzsprung-Russell diagram of clusters. ", "conclusions": "\\begin{figure} \\ifpdf \\includegraphics[width=\\textwidth]{glebbeek_fig_hrd} \\else \\includegraphics[angle=270,width=\\textwidth]{glebbeek_fig_hrd} \\fi \\caption{Colour-magnitude diagram of the open cluster M67 ($\\blacklozenge$). Overplotted are the locations of our models at $4 \\mathrm{Gyr}$, the age of M67. The black ({\\Large$\\bullet$, $\\blacktriangle$}) symbols are collisions from the M67 simulation. Two of these are double collisions, which are indicated by {\\Large$\\blacktriangle$}. The grey ({\\Large$\\bullet$}) symbols are from our larger grid.} \\label{fig:hrd_m67} \\end{figure} Compared to normal stars, collision products are helium enhanced. Most of the helium enhancement is in the interior and does not affect the opacity of the envelope. The helium enhancement does increase the mean molecular weight and therefore the luminosity of the star. This decreases the remaining lifetime of collision products compared to normal stars of the same mass and can be important for the predicted number of blue stragglers from cluster simulations. The increased luminosity changes the distribution of blue stragglers in the colour-magnitude diagram, moving it above the extension of the main sequence. The evolution track of a fully mixed model can be significantly bluer than a self-consistently calculated evolution track of a merger remnant. Fully mixed models are closer to the zero age main sequence. Our grid of models covers most of the observed blue straggler region of M67 (Figure \\ref{fig:hrd_m67}). A better coverage of the blue part of the region can be achieved by increasing the upper mass limit in the grid. The brightest observed blue straggler falls outside the region of our grid because it requires at least a double collision to explain." }, "0710/0710.1506_arXiv.txt": { "abstract": "We have modeled the emission of \\hdo\\ rotational lines from the extreme C-rich star IRC+10216. Our treatment of the excitation of \\hdo\\ emissions takes into account the excitation of \\hdo\\ both through collisions, and through the pumping of the $\\nu_2$ and $\\nu_3$ vibrational states by dust emission and subsequent decay to the ground state. Regardless of the spatial distribution of the water molecules, the \\hdo\\ $1_{10}-1_{01}$ line at 557 GHz observed by the {\\em Submillimeter Wave Astronomy Satellite} (SWAS) is found to be pumped primarily through the absorption of dust-emitted photons at 6 $\\mu$m in the $\\nu_2$ band. As noted by previous authors, the inclusion of radiative pumping lowers the ortho-\\hdo\\ abundance required to account for the 557 GHz emission, which is found to be $(0.5-1)\\times10^{-7}$ if the presence of \\hdo\\ is a consequence of vaporization of orbiting comets or Fischer-Tropsch catalysis. Predictions for other submillimeter \\hdo\\ lines that can be observed by the {\\em Herschel Space Observatory} (HSO) are reported. Multitransition HSO observations promise to reveal the spatial distribution of the circumstellar water vapor, discriminating among the several hypotheses that have been proposed for the origin of the \\hdo\\ vapor in the envelope of IRC+10216. We also show that, for observations with HSO, the \\hdo\\ $1_{10}-1_{01}$ 557 GHz line affords the greatest sensitivity in searching for \\hdo\\ in other C-rich AGB stars. ", "introduction": "\\label{sec:intro} The discovery of water vapor emission in the extreme C-rich AGB star IRC+10216 with SWAS \\citep{mel01}, its confirmation by ODIN \\citep{has06}, and the subsequent detection of other O-bearing molecules like OH \\citep{for03}, H$_2$CO \\citep{for04}, and C$_3$O \\citep{ten06}, have challenged our current understanding of the chemistry in envelopes around C-rich AGB stars. According to standard models, essentially all oxygen nuclei are predicted to be locked into CO or SiO with no reservoir to form other O-bearing molecules (except for low abundances of species such as HCO$^+$ in the outer envelope where photochemistry is important). Therefore, the unexpectedly high abundances found for H$_2$O, OH, and H$_2$CO, indicate that several processes not included in standard models for C-rich environments are driving the oxygen chemistry, but the dominant water production mechanism is still a source of debate. Four distinct mechanisms have been considered as possible sources of the observed water vapor in IRC+10216, each one making a specific prediction for the \\hdo\\ spatial distribution in the envelope: $(i)$ chemistry in the inner envelope, which would imply the presence of \\hdo\\ in the warmest and densest regions; $(ii)$ vaporization of icy orbiting bodies that have survived from the main sequence into the C-rich AGB phase \\citep{mel01,for01}, predicting the release of \\hdo\\ at intermediate radii of $R_{int}=(1-5)\\times10^{15}$ cm; $(iii)$ grain surface reactions, such as the Fischer-Tropsch catalysis on the surfaces of small metallic grains \\citep{wil04}, which predicts \\hdo\\ to attain a nearly uniform abundance at radii larger than $R_{int}=(1.5-2)\\times10^{15}$ cm; $(iv)$ chemistry involving photodissociation products in the outermost regions of the envelope: a specific mechanism relying on the radiative association O+H$_2$$\\rightarrow$\\hdo+$\\gamma$ has been proposed by Ag\\'undez \\& Cernicharo (2006, hereafter AC06), and predicts \\hdo\\ to be present at radii higher than $R_{int}\\approx4\\times10^{16}$ cm. In all cases, \\hdo\\ is expected to have a uniform abundance from $R_{int}$ up to the external region where it is photodissociated producing OH. The only water line detected so far in IRC+10216, the ortho-\\hdo\\ \\t110101\\ transition at 557 GHz, is the ground-state transition, with the upper level at only 27 K above the ground rotational level, and cannot discriminate -by itself- between the processes listed above. Nevertheless, the launch of the Herschel Space Observatory (HSO) will allow us to observe other \\hdo\\ lines in the submillimeter and far-infrared domains, permitting us to infer the region where the \\hdo\\ emission arises, and will potentially favor one of the proposed formation mechanisms. The Heterodyne Instrument for the Far Infrared (HIFI) onboard HSO will provide very high spectral resolution observations (0.14-1.0 MHz), thus permitting the lines to be velocity resolved. In this paper, we model the \\hdo\\ emission from IRC+10216 to show how the \\hdo\\ spatial distribution can be inferred from HSO multi-transition observations. Also, we explore which \\hdo\\ transition provides the most sensitive means of searching for water vapor around AGB stars other than IRC+10216; such a search would determine whether the occurence of \\hdo\\ in C-rich environments is widespread. ", "conclusions": "\\label{sec:discussion} \\subsection{Model uncertainties} Our models for IRC+10216 predict that, whatever the region where \\hdo\\ is formed or released to the outflow, \\hdo\\ is radiatively excited, which implies that model results are independent of the gas temperature and density profiles. This result relies on the extrapolation to higher $T_k$ of the collisional rates given by \\cite{phi96}, which will require further confirmation from new estimates of \\hdo-H$_2$ collisional rates at higher temperatures. Nevertheless, we find that collisional rates would have to be about one order of magnitude higher than estimated to compete efficiently with the radiative rates in the innermost regions of the envelope, which seems somewhat implausible. A relatively uncertain parameter in our models is the assumed distance to IRC+10216, $D=170$ pc. It has been proposed that the distance may be substantially smaller, $D=100-150$ pc \\citep*{zuc86,gro92}. At 170 pc, the inferred stellar luminosity is $L_*\\approx2\\times10^4$ \\Lsun, which results in $L_*\\approx10^4$ \\Lsun\\ at $D=120$ pc, a value more similar to the inferred luminosities of other C-rich AGB stars with similar \\Mdot\\ \\citep*{sch06}. If $D=120$ pc, both \\Mdot\\ and the radiation density at 6 $\\mu$m are a factor of 2 lower than assumed, so that the radiative-to-collisional pumping rate ratio still remains unchanged. Since the closer proximity and weaker 6 $\\mu$m radiation density have opposite effects on the \\hdo\\ outflow rate required to match the SWAS 557 GHz flux, the latter will decrease by less than a factor 2, and thus the expected \\hdo\\ abundance will be a factor of $<2$ higher than in Table~\\ref{tab:h2omodels}. Therefore, we conservatively estimate an o-\\hdo\\ abundance in the range $(0.5-1)\\times10^{-7}$ for $R_{int}=2\\times10^{15}$ cm. Concerning the line ratios to be observed by HSO, we expect values similar to those obtained at $D=170$ pc if the line emission remains unresolved, i.e. for low values of $R_{int}$. For high values of $R_{int}$, beam effects will be more important at $D=120$ pc, diminishing the low-excitation line fluxes relative to the 557 GHz line. A potentially more important source of uncertainty is the assumption of spherical symmetry in our models. Both the molecular and dust emission from IRC+10216 show evidence for departures from a smooth distribution in a spherically symmetric outflow; incomplete, discrete shells or arcs and clumpy structures are instead observed on a wide range of distances to the star \\citep*[e.g.][]{luc99,mau99,fon03,lea06}. Even more important, the OH line shapes observed in IRC+10216 strongly suggest an asymmetric distribution of OH in the outer regions of the envelope \\citep{for03}, thus also suggesting an asymmetric distribution of the parent \\hdo\\ molecule. These asymmetries may alter to some extent the \\hdo\\ line flux ratios calculated with the use of our spherically symmetric approach. While in spherical symmetry the emission from any line is isotropic, in filamentary structures or slabs of velocity-coherent gas the optically thick lines radiate preferentially in the direction perpendicular to the two faces of the sheet \\citep*{eli89}, whereas the emission from thinner lines will approach a more isotropic behavior. Future HSO observations will show the importance of the observed morphological complexity on the line fluxes by showing whether the different \\hdo\\ line flux ratios are consistent with a single value of $R_{int}$, or indicate a range of values. Finally, the models assume a water shell with uniform \\hdo\\ abundance and sharp inner and outer radii; however, a finite abundance gradient obviously takes place at both the \\hdo\\ formation and dissociation regions, and variations of the \\hdo\\ abundance across the shell are also possible. These effects may also alter to some extent the expected line flux ratios. Uncertainties in the outer radius of the \\hdo\\ shell may also affect the expected fluxes from stars with low mass loss rates. \\subsection{Water formation at inner or intermediate radii} Both the cometary and Fischer-Tropsch catalysis hypothesis predict \\hdo\\ to be released or formed at intermediate radii of {\\it a few} $\\times10^{15}$ cm, and although no specific mechanism has been proposed for \\hdo\\ formation at the innermost regions ($R_{int}<10^{15}$ cm), this possibility cannot be rejected. The models shown in section~\\ref{sec:predictions} will permit us to discriminate, from HSO observations of IRC+10216, if there are significant amounts of \\hdo\\ in the innermost regions through the observation of mid-excitation \\hdo\\ transitions. More difficult will be a priori to discriminate between the cometary and Fischer-Tropsch catalysis propositions. Both predict similar values for $R_{int}$; nevertheless, if $R_{int}$ were found to be significantly higher than $2\\times10^{15}$ cm, the release of \\hdo\\ from comets could be favoured, unless some additional mechanism were found to shift $R_{int}$, within the Fischer-Tropsch catalysis framework \\citep{wil04}, outwards in the envelope. The search for water emission at 557 GHz in C-rich AGB stars other than IRC+10216 will also favour one of the two hypotheses: since IRC+10216 is at the high end of mass loss rates from C-rich stars, and the efficiency of Fischer-Tropsch catalysis to form \\hdo\\ molecules is expected to decrease with diminishing \\Mdot, one would expect in such a case a rate of detection significantly lower than that quantified in section~\\ref{sec:crich} for the cometary hypothesis, and one would expect a pronounced decline of the \\hdo\\ abundance with diminishing \\Mdot. The ISO/LWS spectrum of IRC+10216 \\citep{cer96} shows an emission feature at 179.5 $\\mu$m, coincident with the wavelength of the \\t212101\\ o-\\hdo\\ transition, with a flux of $\\approx8\\times10^{-20}$ W cm$^{-2}$. This flux is very similar to that computed for the quoted line in model $B$ ($R_{int}=2.1\\times10^{15}$ cm). However, the far-infrared spectrum of IRC+10216 shows vibrationally excited rotational emission of HCN, and the combined emission of the $\\nu_1=1$ $J=19\\rightarrow18$ and $\\nu_3=1$ $J=19\\rightarrow18$ HCN lines, both emitting at 179.5 $\\mu$m, is expected to be also comparable to the measured line flux at 179.5 $\\mu$m \\citep{cer96}. Given the uncertainties inherent to the HCN model in \\cite{cer96}, where the excited vibrational states are assumed to be thermalized, and given the high density of spectral lines in the IRC+10216 far-infrared spectrum, which makes it difficult to establish the contribution from the $\\nu_1=1$ and $\\nu_3=1$ rotational lines to the spectrum, the relative contribution from HCN and \\hdo\\ to the observed spectral feature is quite uncertain. On the other hand, the expected flux of the \\t303212\\ line at 174.6 $\\mu$m in model $B$ is $\\approx4\\times10^{-20}$ W cm$^{-2}$, below the 3-$\\sigma$ upper limit derived for that line from the ISO/LWS spectrum. These considerations suggest weakly that \\hdo\\ in IRC+10216 is formed or released at radial distances $R_{int}\\gtrsim2\\times10^{15}$ cm. Our models show that the required \\hdo\\ abundance for $R_{int}$ in the range $(2-5)\\times10^{15}$ cm is $(0.5-1)\\times10^{-7}$ relative to H$_2$, which makes the requirements previously demanded for the cometary and Fischer-Tropsch catalysis hypothesis to work in IRC+10216 less restrictive \\citep{for01,wil04}. The water outflow rate is $(0.5-1)\\times10^{-5}$ M$_{\\oplus}$ yr$^{-1}$, which yields, within the framework of vaporization of icy bodies, a required total initial ice mass of $(0.5-10)$ M$_{\\oplus}$ for \\Mdot(\\hdo)/$M_0({\\rm ice})$ in the range $10^{-5}-10^{-6}$ yr$^{-1}$ \\citep{for01}. On the other hand, the Fischer-Tropsch catalysis on metallic grains will require a density of iron grains relative to total H nuclei of $n_g({\\rm Fe})/n_{{\\rm H}}=(0.5-1)\\times10^{-13}$ to explain the observed \\hdo\\ 557 GHz emission \\citep{wil04}. \\subsection{Water formation in the outermost layers} \\label{sec:radassoc} Multitransition HSO observations of \\hdo\\ in IRC+10216 will easily establish whether or not water is formed in the external layers of the envelope; the predicted fluxes in Fig.~\\ref{fig:flujosrint2} and the HSO-HIFI sensitivities in Table~\\ref{tab:h2otrans} indicate that only the \\t110101\\ o-\\hdo\\ line, and possibly the \\t111000\\ and \\t202111\\ p-\\hdo\\ lines, are detectable in one hour of observing time for $R_{int}\\gtrsim4\\times10^{16}$ cm. For high $R_{int}$, the expected flux in the 557 GHz line can be obtained analytically. Since essentially all o-\\hdo\\ molecules are in the ground $1_{01}$ level (Fig.~\\ref{fig:pump557ghz}a), the o-\\hdo\\ \\t110101\\ line flux is derived from the radiative pumping rate given in eq.~(\\ref{gammar}) after multiplying $\\Gamma_r$ by 1.35 to account for the two radiative pumping routes that are relevant at high distances from the star (Fig.~\\ref{fig:pump557ghz}b): \\begin{eqnarray} F(1_{10}\\rightarrow1_{01}) = 1.4\\times10^{-21} \\times \\left( \\frac{4\\times10^{17} \\, {\\rm cm}}{R_{out}} \\right) \\nonumber \\\\ \\times \\left( \\frac{R_{out}}{R_{int}} -1 \\right) \\times \\left( \\frac{X({\\rm o-H_2O})}{5\\times10^{-7}} \\right) \\,\\, {\\rm \\frac{W}{cm^{2}}}, \\label{fluxanal} \\end{eqnarray} where $X$ is the abundance relative to H$_2$. Equation~(\\ref{fluxanal}) is independent of the assumed distance $D$ to the star because, in order to match the observed mid-IR continuum, $\\Gamma_r\\propto D^{2}$ (eq.~\\ref{eqj}); it assumes a water shell with sharp edges at $R_{int}$ and $R_{out}$ and constant \\hdo\\ abundance; it ignores slight opacity effects in the ro-vibrational lines (section~\\ref{sec:110101}) as well as beam effects (both slightly raise the required abundance), and overestimates by less than 20\\% the 557 GHz line flux obtained in model $C$. In the limit $R_{out}\\rightarrow\\infty$, and using $F(1_{10}\\rightarrow1_{01})=10^{-20}$ W cm$^{-2}$ \\citep{mel01}, eq.~(\\ref{fluxanal}) gives \\begin{equation} \\chi({\\rm H_2O}) \\gtrapprox 2.4\\times10^{-7} \\times \\left( \\frac{R_{int}}{4\\times10^{16} \\, {\\rm cm}} \\right), \\label{fluxanallim} \\end{equation} where $\\chi({\\rm H_2O})$ is here the \\hdo\\ (ortho+para) abundance relative to H nuclei. Equation~\\ref{fluxanallim} gives the sharp-inner edge, lower limit for the \\hdo\\ abundance required to account for the observed 557 GHz line flux, as a function of $R_{int}$. AC06 have reported an abundance of $\\chi({\\rm H_2O})\\sim10^{-7}$ to account for the observed 557 GHz line flux in IRC+10216. However, the quoted abundance would imply an inner radius of $R_{int}\\lessapprox2\\times10^{16}$ cm, but the \\hdo\\ abundance shown by AC06 (their Fig. 7) decreases sharply at $r<4\\times10^{16}$ cm. Based on detailed modelling, we indicate that the $\\chi({\\rm H_2O})$ profile given by AC06 has to be shifted up by a factor of $\\approx2.5$ to account more accurately for the 557 GHz line flux measured by SWAS. In the model proposed by AC06, atomic oxygen is produced in a shell by the photodissociation of CO (and particularly the $^{13}$CO isotopologue, which shields itself far less effectively than $^{12}$CO.) Since the temperature is low ($\\sim 10$~K) within the shell where the atomic oxygen abundance is significant, the neutral-neutral reaction sequence $$\\rm O + H_2 \\rightarrow OH + H$$ $$\\rm OH + H_2 \\rightarrow H_2O + H$$ is very slow --and therefore is negligible as a source of H$_2$O-- the first reaction being endothermic and the second --although exothermic-- possessing a substantial activation energy barrier. AC06 therefore proposed the radiative association reaction $$\\rm O + H_2 \\rightarrow H_2O + \\gamma$$ as an alternative source of H$_2$O. To match the SWAS- and ODIN-observed water line fluxes, AC06 had to posit that this reaction is relatively rapid at low temperature, with a rate coefficient $k_{ra} \\rm \\sim 10^{-15}\\,cm^3\\,s^{-1}$. According to our estimation above, $k_{ra}$ has to be a factor $\\approx2.5$ higher than this value to account for the measured \\t110101\\ \\hdo\\ flux. Nevertheless, and even with the use of the $k_{ra}$ estimation given by AC06, we find that this large reaction rate coefficient for the radiative association of O and H$_2$ is inconsistent with observations of H$_2$O and OH in at least one translucent molecular cloud, for which sufficient data exist to disipate any significant ambiguity: observations of the cloud along the sight-line to HD 154368 --carried out by \\cite{spa98} with the use of the Goddard High Resolution Spectrograph (GHRS) on the {\\it Hubble Space Telescope} (HST) -- yield a 3~$\\sigma$ upper limit on the water abundance that lies almost two orders of magnitude below the value that would obtain were $k_{ra}$ as large as $\\rm 10^{-15}\\,cm^3\\,s^{\\rm -1}$. The factor of discrepancy would rise above $150$ for $k_{ra} \\rm \\approx 2.5 \\times 10^{-15}\\,cm^3\\,s^{-1}$. The best-fit model for the HD 154368 cloud obtained by \\cite{spa98} posits a plane parallel cloud of total visual extinction $A_V = 2.65$~mag in which the density of H nuclei is $n_{\\rm H} = 325\\, \\rm cm^{-3}$ \\citep[this is probably a lower limit at the cloud center, as the density inferred from the CO $J=1\\rightarrow0/3\\rightarrow2$ ratio is $\\sim10^3$ cm$^{-3}$; see][]{van91} and the external ultraviolet radiation field is 3 times mean interstellar value given by \\cite{dra78}: $I_{UV} = 3$. Under these conditions, the destruction of H$_2$O is dominated by photodissociation at a rate $\\zeta = 5.9 \\times 10^{-10}{\\rm s}^{-1} \\times I_{UV} \\times (\\exp\\{-1.7 A_{V1}\\} + \\exp\\{-1.7 A_{V2}\\})/2$ \\citep*{let00}\\footnote{We note that the factor of 1.7 in the exponential, widely used in the literature, yields \\hdo\\ photodissociation rates at the midplane of the plane-parallel cloud that are higher than the values reported by \\cite{rob91} by factors of 2.2 and 10 for $A_V^{tot}=1$ mag and $A_V^{tot}=10$ mag, respectively; therefore, our estimation for the $N({\\rm H_2O})/N({\\rm O})$ ratio predicted by the radiative association of O and H$_2$ is probably a conservative lower limit.}, where $A_{V1}$ is the visual extinction to one cloud surface and $A_{V2} = 2.65 - A_{V1}$ is the extinction to the other. At the cloud center, the water photodissociation rate is $1.9 \\times 10^{-10}\\,\\,{\\rm s}^{-1}$. For a radiative association rate of $k_{ra}$, the ratio of water vapor to atomic oxygen is therefore given by the expression $$n({\\rm H_2O})/n({\\rm O}) = k_{ra} n_{\\rm H} f_{\\rm H_2}/ \\zeta,$$ where $f_{\\rm H_2} \\equiv n({\\rm H_2}) / n_{\\rm H}$. At the cloud center, the cloud is almost fully molecular, with $f_{\\rm H_2} \\sim 0.5$, and the above equation yields $n({\\rm H_2O})/n({\\rm O}) = 8.7 \\times 10^{-4} \\,(k_{ra}/\\rm 10^{-15}\\,cm^3\\,s^{\\rm -1})$. In Fig.~\\ref{fig:nh2ono}, we show the predicted $n({\\rm H_2}) / n_{\\rm H}$ ratio for the best-fit \\cite{spa98} model, together with the $n({\\rm H_2O})/n({\\rm O})$ ratio that would result if $k_{ra}$ were $\\rm 10^{-15}\\,cm^3\\,s^{-1}$. The model predicts hydrogen to be predominantly in molecular form, in agreement with results by \\cite{sno96}. Averaging the $n({\\rm H_2O})/n({\\rm O})$ ratio over the entire sight-line, we obtain a column density ratio $N({\\rm H_2O})/N({\\rm O}) = 5.3 \\times 10^{-4} \\,(k_{ra}/\\rm 10^{-15}\\,cm^3\\,s^{\\rm -1})$. The atomic oxygen column density along the HD 154368 sight-line is $N({\\rm O}) = 1.2 \\times 10^{18}\\,\\,{\\rm cm^{-2}}$ (Snow et al. 1996, from absorption line observations of the 1355 \\AA\\ OI] line). Based on a search for the $C^1B_1 - X^1A_1$ band of water vapor near 1240 \\AA, \\cite{spa98} obtained 3~$\\sigma$ upper limit on the water column density of $9 \\times 10^{12}\\,\\,{\\rm cm^{-2}}$, corresponding to $N({\\rm H_2O})/N({\\rm O}) = 7.5 \\times 10^{-6}$. This 3~$\\sigma$ upper limit lies a factor 70 below the value that we would obtain were $k_{ra}$ equal to $10^{-15}\\rm \\,cm^3\\,s^{-1}$ as AC06 suggested, and places a 3~$\\sigma$ upper limit of $1.4 \\times 10^{-17}\\rm \\,cm^3\\,s^{-1}$ on $k_{ra}$. AC06 suggested that the freeze out of oxygen onto grain mantles could diminish the water vapor abundance in molecular clouds, as proposed by \\cite{ber00} to explain the low H$_2$O abundances measured by SWAS in dense clouds. Ice absorption line observations of diffuse/translucent sight-lines, however, indicate that water ice is generally present only in clouds of $A_V \\ge 3$ \\citep{whi01}. Furthermore, in the specific case of HD 154368 under present consideration, the atomic oxygen is known from direct measurement to be $1.2 \\times 10^{18}\\,\\,{\\rm cm^{-2}}$. The column density of H nuclei along this sight-line, $N_{\\rm H} = N({\\rm H}) + 2 N({\\rm H}_2)$, has been measured to be $4.2 \\times 10^{21}\\,\\,\\rm cm^{-2}$ \\citep{sno96}, so the mean line-of-sight oxygen abundance is $n({\\rm O})/n_{{\\rm H}} \\sim 3 \\times 10^{-4}$, a value that is entirely consistent with the abundances measured along diffuse sight-lines \\citep{mey98} and inconsistent with a significant depletion of oxygen onto ice mantles. In summary, the upper limit on the water vapor abundance observed towards HD 154368 definitively rules out a rate coefficient for the radiative association reaction $\\rm O + H_2 \\rightarrow H_2O + \\gamma$ that is large enough to explain --in the context of the AC06 model-- the H$_2$O \\t110101\\ line strength measured by SWAS toward IRC+10216. If HSO observations would indicate that \\hdo\\ is formed in the external layers of IRC+10216, an explanation other than the radiative association proposed by AC06 would be required to avoid incompatibilities with observations toward HD 154368." }, "0710/0710.4786_arXiv.txt": { "abstract": "The proper motion measurements for 143 previously known L and T dwarfs are presented. From this sample we identify and discuss 8 high velocity L dwarfs. We also find 4 new wide common proper motion binaries/multiple systems. Using the moving cluster methods we have also identified a number of L dwarfs that may be members of the Ursa Major (age $\\approx$ 400 Myr), the Hyades (age $\\approx$ 625 Myr) and the Pleiades (age $\\approx$ 125 Myr) moving groups. ", "introduction": "Brown dwarfs may be thought of as failed stars. These low mass ($\\leq$70 M$_{\\rm Jup}$ \\citet{burrows01}), cool objects are the lowest mass objects that the star formation process can produce. The majority of the brown dwarfs that have been discovered to date are field objects discovered using surveys such as the Two Micron All Sky Survey (2MASS; \\citet{skrutskie06}, see \\citet{leggett02} for examples), the DEep Near-Infrared Sky survey (DENIS; \\citet{denis05}, see \\citet{delfosse99} for examples), the Sloan Digital Sky Survey (SDSS;\\citet{york00} see \\citet{hawley02} for examples) and the UKIRT Deep Infrared Sky Survey (UKIDSS; \\citet{lawrence06}, see \\citet{kendall07} for examples). However, to study brown dwarfs in depth, a knowledge of their age is essential, which means we must study brown dwarfs in open star clusters or moving groups. Once a brown dwarf has been proved to belong to an open star cluster, or a moving group, then the age of the dwarf is known, allowing meaningful comparisons to evolutionary models to be made. The most recent example of this is the study done by \\citet{bannister07} who used existing proper motions and parallax measurements to show that a selection of field dwarfs in fact belong to the Ursa Major and Hyades moving groups. The importance of this study, is that these are the first brown dwarfs to be associated with an older cluster or group. Older clusters such as the Hyades are expected to contain very few or no brown dwarfs or low mass members, due to the dynamical evolution of the cluster over time \\citep{adams02}. However, these escaped low mass objects may remain members of the much larger moving group that surrounds the cluster. To continue the study started by \\citet{bannister07}, proper motions need to be measured for the majority of the field brown dwarfs currently known. This has been accomplished using the wide field camera (WFCAM, \\citet{casali07}) of the United Kingdom Infrared Telescope (UKIRT). Using these WFCAM images and existing data we have measured proper motions for 143 L and T dwarfs listed in the online Dwarf Archive\\footnote{See http://spider.ipac.caltech.edu/staff/davy/ARCHIVE/, a webpage dedicated to L and T dwarfs maintained by C.\\ Gelino, D.\\ Kirkpatrick, and A.\\ Burgasser.}. This proper motion data may be put to a number of uses. Taken with measured radial velocities and distances, it can yield all three components of velocity (U,V,W). Using reduced proper motion diagrams it can be used as an approximate measure of distance. However we have no radial velocities for these objects. These proper motion measurements can however also be used to help identify objects as members of a star cluster or members of a moving group via the moving cluster method. Our proper motion data is discussed and listed in section 2 of this paper. From the proper motion measurements, we find 5 new wide common proper motion binaries/multiple systems. We also identify 8 high velocity L dwarfs, which are discussed in section 4. We suggest that these L dwarfs are probably old and belong to the thick disc or halo population of the galaxy. This in turn suggests that they are likely to be very faint stars rather than brown dwarfs. Finally in section 5 we identify a number of L and T dwarfs that may be members of the Hyades, Pleiades and Ursa Major moving groups . ", "conclusions": "We have measured the proper motions for 143 dwarfs from the Dwarf Archive. From these measurements, we find 4 new common proper motion wide binary or multiple systems. We also identify 8 high velocity dwarfs i.e. dwarfs with tangential velocities $\\geq$ 100 kms$^{-1}$. These dwarfs also have bluer than average \\textit{J}-\\textit{K} colours. We argue that these are probably thick disc objects with an age of order 10 Gyr. We estimate their luminosities which are $\\approx$10$^{-4}$L/L$_{\\odot}$. This suggests that they are probably very low luminosity stars rather than brown dwarfs. If so, they may be some of the dimmest stars found to date. Finally we have found 15 L dwarfs that are potential members of the Hyades moving group, 5 that are potential members of the Ursa Major moving group and 5 that are potential members of the Pleiades moving group. The next obvious step towards confirming membership of these groups is to measure parallaxes for these dwarfs. Parallaxes will allow accurate distances to be used to compare with the moving group distance. \\textit{Spitzer} 3.5, 4.49, 5.73 and 7.87 micron magnitudes will also be valuable for a fuller understanding of the high velocity metal poor dwarfs." }, "0710/0710.3359_arXiv.txt": { "abstract": "The feasibility of a mean-field dynamo in nonhelical turbulence with superimposed linear shear is studied numerically in elongated shearing boxes. Exponential growth of magnetic field at scales much larger than the outer scale of the turbulence is found. The charateristic scale of the field is $\\lB\\propto S^{-1/2}$ and growth rate is $\\gamma\\propto S$, where $S$ is the shearing rate. This newly discovered shear dynamo effect potentially represents a very generic mechanism for generating large-scale magnetic fields in a broad class of astrophysical systems with spatially coherent mean flows. ", "introduction": " ", "conclusions": "" }, "0710/0710.1440_arXiv.txt": { "abstract": "% The UV excess shown by elliptical galaxies in their spectra is believed to be caused by evolved low-mass stars, in particular sdB stars. The stellar system most similar to the ellipticals for age and metallicity, in which it is possible to resolve these stars, is the bulge of our Galaxy. sdB star candidates were observed in the color magnitude diagram of a bulge region by Zoccali et al.\\ (2003). The follow-up spectroscopic analysis of these stars confirmed that most of these stars are bulge sdBs, while some candidates turned out to be disk sdBs or cool stars. Both spectroscopic and photometric data and a spectral library are used to construct the integrated spectrum of the observed bulge region from the UV to the optical: the stars in the color magnitude diagram are associated to the library spectra, on the basis of their evolutionary status and temperature. The total integrated spectrum is obtained as the sum of the spectra associated to the color magnitude diagram. The comparison of the obtained integrated spectrum with old single stellar population synthetic spectra calculated by Bruzual \\& Charlot~(2003) agrees with age and metallicity of the bulge found by previous work. The bulge integrated spectrum shows only a very weak UV excess, but a too strict selection of the sample of the sdB star candidates in the color magnitude diagram and the exclusion of post-Asymptotic Giant Branch stars could have influenced the result. ", "introduction": "\\label{sec:intro} The UV excess that elliptical galaxies and bulge of spiral galaxies show in their spectra at $\\lambda$ shorter than 2300~\\AA~ was one of the most puzzling discoveries in the last 30 years, since these stellar systems are believed to be old and metal rich, without young and massive stars emitting most of their flux at short wavelength. It is now widely accepted that this UV emission is caused by evolved low mass stars, in particular Extreme Horizontal Branch stars (EHB), called also sdB stars from their spectral classification. These stars are faint in the optical wavelength range and with the current instrumentation it is impossible to resolve them in the nearest galaxies. The stellar system most similar to the ellipticals for age and metallicity in which it is possible to resolve sdB stars is the bulge of our Galaxy. A sample of sdBs star candidates was observed in the Galactic bulge by Zoccali et al.~(2003)\\nocite{zoccali03}, by means of $V$ and $I$ photometry of the region MW05 from the ESO Imaging Survey (EIS\\footnote{\\tt http://www.eso.org/science/eis/}, the observations were taken with the Wide Field Imager, WFI@2.2m). These stars could be either highly reddened sdBs or cooler stars affected by lower reddening. A follow-up spectroscopic analysis of these stars has been necessary and observations at the Very Large Telescope (VLT) telescope were obtained. The data reduction and the comparison of the obtained spectra with models of hot evolved stars confirmed indeed that most of these stars are bulge sdBs, while some candidates turned out to be disk sdBs or cool stars (for more details, see Busso et al.~2005\\nocite{busso05}). To be sure that the observed bulge region was not peculiar, other bulge fields were searched for sdB candidates: EIS photometric data of the bulge fields MW07 and MW08 were reduced and analyzed, finding that sdB star candidates are present also in these fields. This work presents the procedure adopted (following the recipe as in Santos et al.~1995) to construct the integrated spectrum of the bulge region MW05. ", "conclusions": "" }, "0710/0710.5780_arXiv.txt": { "abstract": "We report high-resolution spectroscopy of 125 field stars previously observed as part of the Sloan Digital Sky Survey and its program for Galactic studies, the Sloan Extension for Galactic Understanding and Exploration (SEGUE). These spectra are used to measure radial velocities and to derive atmospheric parameters, which we compare with those reported by the SEGUE Stellar Parameter Pipeline (SSPP). The SSPP obtains estimates of these quantities based on SDSS $ugriz$ photometry and low-resolution ($R \\sim 2000$) spectroscopy. For F- and G-type stars observed with high signal-to-noise ratios ($S/N$), we empirically determine the typical random uncertainties in the radial velocities, effective temperatures, surface gravities, and metallicities delivered by the SSPP to be 2.4 km s$^{-1}$, 130 K (2.2 \\%), 0.21 dex, and 0.11 dex, respectively, with systematic uncertainties of a similar magnitude in the effective temperatures and metallicities. We estimate random errors for lower $S/N$ spectra based on numerical simulations. ", "introduction": "Starting from the sixth public data release (DR-6; Adelman-McCarthy et al. 2007), the Sloan Digital Sky Survey (SDSS) provides estimates of the atmospheric parameters for a subset of the stars observed spectroscopically in the survey (those in the approximate range of temperature $4500 \\le T_{\\rm eff} \\le 7500$~K). Following completion of the main survey (SDSS-I), the SDSS instrumentation has been devoted to several programs, including SEGUE: Sloan Extension for Galactic Understanding and Exploration, a massive survey of the stellar content of the Milky Way. Collectively, the suite of computer programs employed to determine atmospheric parameters from SEGUE data is known as the SEGUE Stellar Parameter Pipeline (SSPP). Because each of the public data releases of the SDSS includes and supersedes previous releases, DR-6 also includes atmospheric parameters for archival stellar observations in SDSS-I. These stellar parameters are derived by a series of methods, some of which consider purely spectroscopic information (continuum-normalized spectra), solely photometry (available in the survey's $ugriz$ system for all targets), or a combination of photometry and spectroscopy. Paper I in this series describes the SSPP in detail (Lee et al. 2007a). Paper II compares the predictions of the SSPP radial velocities and atmospheric parameters with likely members of Galactic globular and open clusters (Lee et al. 2007b). The SDSS uses a CCD camera (Gunn et al. 1998) on a dedicated 2.5m telescope (Gunn et al. 2006) at Apache Point Observatory, New Mexico, to obtain images in five broad optical bands ($ugriz$; Fukugita et al.~1996) over approximately 10,000~deg$^2$ of the high Galactic latitude sky. The survey data-processing software measures the properties of each detected object in the imaging data in all five bands, and determines and applies both astrometric and photometric calibrations (Lupton et al. 2001; Pier et al. 2003; Ivezi\\'c et al.~2004). Photometric calibration is provided by simultaneous observations with a 20-inch telescope at the same site (Hogg et al.~2001; Smith et al.~2002; Stoughton et al.~2002; Tucker et al.~2006). A technical summary is provided by York et al. (2000). SDSS-I and the ongoing SEGUE survey have already built the largest-ever catalog of stars in the Milky Way. To date, this includes photometry in five bands for over 200 million stars and spectroscopy for nearly 300,000 stars (Adelman-McCarthy et al. 2007). The SDSS spectrographs deliver a resolving power $\\lambda/$FWHM $\\sim 2000$ over the wavelength range 380-900 nm. Data reduction is fully automated, and the SSPP employs the final products from the SDSS pipeline as input to produce atmospheric parameters (effective temperature, surface gravity, and metallicity) for stars with spectral types A, F, G, and K. The best results are obtained for F- and G-type stars spanning the effective temperature range $5000 < T_{\\rm eff} < 7000$~K. The quality of the SSPP atmospheric parameters is evaluated using different approaches, as already described in Paper I: comparing with previously published spectral libraries, well-studied open and globular clusters, and with high-resolution observations of field stars. Existing spectral libraries are useful in order to evaluate and calibrate the SSPP methods that rely on spectroscopy alone. Allende Prieto et al. (2006) employed the low-resolution Indo-US library (Valdes et al. 2004), and high-resolution spectra from the Elodie library (Prugniel \\& Soubiran 2001) and the S$^4$N archive (Allende Prieto et al. 2004). Because the $ugriz$ system was introduced with the SDSS, the stars included in existing spectral libraries lack photometry in this system. In addition, these are relatively bright stars, typically with $V<14$ mag, brighter than the bright magnitude limit of the SDSS imaging. The bright magnitude limit for the SDSS is set by the saturation threshold of the detectors at the sidereal driftscan rate of the survey. Obtaining data for these brighter stars would require special-purpose observations with a very different instrument configuration, which would call into question their value as calibration observations for the otherwise homogeneous imaging survey. Star clusters provide stringent tests of the SSPP, as the same metallicity should be derived for stars that explore wide ranges of masses and luminosities. Paper II in this series examines SSPP results for likely members of clusters included in DR-6. One cannot choose clusters with any given metallicity, but has to take what is provided by nature and accessible from Apache Point. Furthermore, the effective temperatures and surface gravities for the members of any given cluster are very strongly correlated, depending on age and chemical composition. This leads to a patchy coverage of the parameter space. Field stars, on the other hand, can be chosen to provide better coverage and, therefore, naturally complement the clusters. Among the stars spectroscopically observed with SDSS, those in the range $14 < V <16.5$ mag can be observed at high spectral resolution with large-aperture telescopes and modest integration times. Due to the vast size of the SDSS stellar sample, these stars can be selected to more uniformly cover the parameter space of stellar properties, and have the additional benefit that photometry is already available for them in the SDSS native system. This paper, the third in the SSPP series, is devoted to the analysis of 125 SDSS stars newly observed at high-resolution with the Hobby-Eberly, Keck, and Subaru telescopes. Section 2 describes the sample selection and the observations. The determination of radial velocities and atmospheric parameters, based on these observations, are discussed in \\S 3 and \\S 4, respectively. Section 5 describes the results for several well-known standard stars observed with the Hobby-Eberly Telescope. Section 6 compares the parameters derived from high-resolution spectroscopy with those from the SSPP. Section 7 describes numerical experiments that explore how the parameters degrade at lower signal-to-noise ratios. Our conclusions are summarized in \\S 8. ", "conclusions": "We report on an analysis of high-resolution spectroscopic observations of a sample of stars previously observed with the SDSS instrumentation as part of SDSS-I or SEGUE. These new data are used to derive radial velocities and atmospheric parameters, and to scrutinize the performance of the SSPP Pipeline described in Paper I in this series. The sample we have examined includes 81 stars observed with the HET-HRS, 25 stars obtained with Keck-ESI, 11 stars observed with Keck-HIRES, and 9 stars from Subaru-HDS. Through a comparison with external spectroscopic libraries, and by employing multiple methods of analysis for the HET sample, we estimate that our reference radial velocities are accurate to 1.6 km s$^{-1}$. Our values for the stellar atmospheric parameters, effective temperature, surface gravity, and metallicity, are accurate to 1.5 \\% ($\\sim 90$~K), 0.13~dex and 0.05~dex, respectively. These figures are derived from the comparison with the parameters for nearby stars in the S$^4$N catalog, but we find they are still valid for the moderately high $S/N$ of the HET spectra. Using the HET sample to benchmark the SSPP, subtracting in quadrature the uncertainties in the results for the former, we conclude that the SDSS/SEGUE radial velocities are typically accurate to 2.4 km s$^{-1}$ for high signal-to-noise SDSS spectra ($S/N > 50/1$). A similar comparison of the atmospheric parameters returned by the SSPP with those obtained from HET spectra leads to the conclusion that the SSPP effective temperatures, surface gravities, and metallicities for bright targets show random errors of 2.2\\% ($\\sim 130$ K), 0.21 dex, and 0.11 dex, respectively. Systematic offsets of a similar size are detected for the effective temperatures and metallicities. We evaluate the expected random uncertainties as a function of $S/N$ by repeating the analysis after introducing noise in the SDSS spectra. More extended tests are underway and will be reported elsewhere. Our study also finds that the internal uncertainties delivered by the SSPP for both radial velocities and atmospheric parameters need to be systematically increased by a factor of $2-3$ in order to be consistent with our derived external errors. The uncertainties in the average SSPP atmospheric parameters are simply derived as the standard error of the mean for a Normal distribution from the multiple techniques applied to any particular target. The fact that many methods share the same spectroscopic indicators (e.g. Balmer lines or SDSS color indices to gauge $T_{\\rm eff}$), and models (e.g. Kurucz's model atmospheres) may cause unaccounted correlations that result in underestimated uncertainties. The validation and calibration of the SSPP is an ongoing project. Several additional open and globular clusters have recently had data obtained with SDSS instrumentation, and will be considered in future papers. A sample of up to several hundred very low-metallicity stars from SDSS/SEGUE is presently being observed with the HET, which we will add to our calibration sample. Additional stars of intermediate metallicity, and with hotter and cooler temperatures than considered in the present work, will be added to our calibration sample based on observations with a number of large-aperture telescopes. Our goal is to produce an SSPP validation catalog for on the order of 500 stars, which will be used to refine and adjust the individual parameter estimation techniques employed by the SSPP, and thus establish a definitive atmospheric parameter estimation scale for application to the large (and growing) SDSS/SEGUE stellar samples, as well as to other future surveys." }, "0710/0710.5922_arXiv.txt": { "abstract": "\\noindent We derive upper limits on the ratio $\\fgrbccsn (z)\\equiv \\rgrb(z) /\\rccsn(z)\\, \\equiv\\fgrbccsn(0)(1+z)^\\alpha$, the ratio of the rate, $\\rgrb$, of long-duration Gamma Ray Bursts (GRBs) to the rate, $\\rccsn$, of core-collapse supernovae (CCSNe) in the Universe ($z$ being the cosmological redshift and $\\alpha\\geq 0$), by using the upper limit on the diffuse TeV--PeV neutrino background given by the AMANDA-II experiment in the South Pole, under the assumption that GRBs are sources of TeV--PeV neutrinos produced from decay of charged pions produced in $p\\gamma$ interaction of protons accelerated to ultrahigh energies at internal shocks within GRB jets. For the assumed ``concordance model'' of cosmic star formation rate, $\\rsf$, with $\\rccsn (z) \\propto \\rsf (z)$, our conservative upper limits are $\\fgrbccsn(0)\\leq 5.0\\times10^{-3}$ for $\\alpha=0$, and $\\fgrbccsn(0)\\leq 1.1\\times10^{-3}$ for $\\alpha=2$, for example. These limits are already comparable to (and, for $\\alpha\\geq 1$, already more restrictive than) the current upper limit on this ratio inferred from other astronomical considerations, thus providing a useful independent probe of and constraint on the CCSN-GRB connection. Non-detection of a diffuse TeV--PeV neutrino background by the up-coming IceCube detector in the South pole after three years of operation, for example, will bring down the upper limit on $\\fgrbccsn (0)$ to below few $\\times10^{-5}$ level, while a detection will confirm the hypothesis of proton acceleration to ultrahigh energies in GRBs and will potentially also yield the true rate of occurrence of these events in the Universe. ", "introduction": "} Detection of supernova (SN) features in the afterglow spectra of several long duration (typically $>2\\s$) Gamma Ray Bursts (GRBs) in the past one decade has provided strong support to the hypothesis that a significant fraction, if not all, of the long duration GRBs arise from collapse of massive stars; see, e.g., Refs.~\\cite{Meszaros_araa02,Woosley-Bloom_araa06,DellaValle06} for recent reviews. The observed SN features in the GRB afterglow spectra are similar to those usually associated with core-collapse supernovae (CCSNe) of Type Ib/c (see, e.g., ~\\cite{Woosley-Heger06,DellaValle06}). The total energy (corrected for beaming) in keV--MeV gamma rays emitted by typical long-duration GRBs is of order $10^{51}\\erg$, which is roughly the same as the total explosion energy seen in typical CCSNe, although there exists considerable diversity in the energetics of both the SN and the GRB components in the SN-GRB associations observed so far. In particular, the estimated explosion energies of the SNe associated with the GRBs observed so far seem to be somewhat larger than those of normal SNe, leading to this ``special'' class of SNe being sometimes referred to as ``hypernovae''. The broad class of observational results on SN-GRB associations can be understood within the context of the ``collapsar'' model~\\cite{collapsar_model_refs} in terms of a simple phenomenological picture (see, e.g., ~\\cite{Woosley-Zhang_astroph_0701320}) in which the core-collapse of a massive Wolf-Rayet star gives rise to two kinds of outflows emanating from the central regions inside the collapsed star: (a) a narrowly collimated and highly relativistic jet that is responsible for the GRB activity, the jet being driven, for example, by a rapidly rotating and accreting black hole formed at the center in the core-collapse process, and (b) a more wide-angled, quasi-spherical and non-relativistic (or at best sub-relativistic) outflow that goes to blow up the star and gives rise to the supernova. The energies channeled into these two components may in general vary independently, which may explain the diversity of energetics in the observed SN-GRB associations. Actually, depending on the energy contained in it the ``GRB jet'' may or may not be able to penetrate through the stellar material and emerge outside. Indeed, the fact that the SN-GRB associations observed so far involve CCSNe of Type Ib/c, but not of Type II, may be due to the inability of the GRB-causing jet to penetrate through the relatively larger amount of outer stellar material in the case of Type II SN as compared to that in SNe of Type Ib/c~\\cite{fn1}. Considering various factors that may govern the energy channeled into the GRB-causing jet, such as the mass and rotation rate of the black hole, accretion efficiency, efficiency of conversion of accretion energy into collimated relativistic outflow, and so on, Woosley and Zhang~\\cite{Woosley-Zhang_astroph_0701320} have obtained a rough lower limit of $\\sim10^{48}\\erg/\\s$ for the power required for the jet to be able to emerge from the star. This is consistent with the energetics of the GRB components of the SN-GRB associations observed so far. While SN-GRB associations strongly support the stellar core-collapse origin of most long-duration GRBs, clearly, not all core-collapse events may result in a GRB --- the latter depends on whether or not the core-collapse event actually results in a ``central engine'' (a rotating black hole fed by an accretion disk in the above mentioned phenomenological picture, for example) that is capable of driving the required collimated relativistic outflow. In other words, while every long-duration GRB would be expected to be accompanied by a core-collapse supernova~\\cite{fn2}, the reverse is not true in general. What fraction of all stellar core-collapse events in the universe produce GRBs? Methods based on astronomical observations generally indicate the ratio between the cosmic GRB rate and the cosmic Type Ib/c SN rate, $\\fgrbsnIbc$, to be in the range $\\sim 10^{-3}$ -- $10^{-2}$ for a wide variety of different assumptions on various relevant parameters such as those that characterize the cosmic star formation rate (SFR), initial mass function (IMF) of stars, masses of Type Ib/c SN progenitors, the luminosity function of GRBs, the beaming factor of GRBs (associated with the fact that individual GRB emissions are highly non-isotropic and confined to narrowly collimated jets covering only a small fraction of the sky), and so on; see, for example, \\cite{Guetta-DellaValle_06,Bissaldi_etal_07} and references therein. The dominant uncertainty in the estimate of $\\fgrbsnIbc$ comes from the uncertainties in the estimates of the local GRB rate and the average GRB beaming factor. However, irrespective of the exact value of the ratio $\\fgrbsnIbc$, it is clear that this ratio is significantly less than unity. This indicates that, apart from just being sufficiently massive stars, the GRB progenitors may need to satisfy additional special conditions. For example, it has been suggested~\\cite{Woosley-Bloom_araa06} that the degree of rotation of the central iron core of the collapsing star and the metalicity of the progenitor star may play crucial roles in producing a GRB. In this paper, we discuss an alternative probe of the cosmic GRB rate that uses the predicted high energy (TeV--PeV) diffuse neutrino background produced by GRBs and the experimental upper limit on high energy diffuse neutrino background given by the AMANDA-II experiment in the South Pole~\\cite{amanda_II_limit}. Existence of a high (TeV--PeV) energy diffuse GRB neutrino background (DGRBNuB) due to $p\\gamma$ interactions of (ultra)high energy protons accelerated within GRB sources is a generic prediction~\\cite{Waxman-Bahcall_grbnu} in most currently popular models of GRBs. This DGRBNuB is subject to being probed by the currently operating and up-coming large volume (kilometer scale) neutrino detectors such as IceCube~\\cite{icecube}, ANITA~\\cite{anita}, ANTARES~\\cite{antares}, for example. Since neutrinos, unlike electromagnetic radiation, can travel un-hindered from the furthest cosmological distances, the DGRBNuB automatically includes the contributions from all GRBs in the Universe. Thus, an analysis of the DGRBNuB is likely to provide a good picture of the true rate of occurrence of these events in the Universe. Indeed, as we show in this paper, the upper limits on $\\fgrbccsn (0)$, the ratio of the local (i.e., redshift $z=0$) GRB to CCSN rates, derived here from the consideration of DGRBNuB, are, for a wide range of values of the relevant parameters, already more restrictive than the current upper limit on this ratio ($\\sim 2.5\\times10^{-3}$) inferred from other astronomical considerations~\\cite{Guetta-DellaValle_06,Bissaldi_etal_07}. Further, non-detection of a diffuse TeV--PeV neutrino background by the up-coming IceCube detector~\\cite{icecube} in the South Pole after three years of operation, for example, will imply upper limits on $\\fgrbccsn(0)$ at the level of few \\ $\\times10^{-5}$, while a detection of the DGRBNuB will provide strong support to the hypothesis of proton acceleration to ultrahigh energies within GRB jets. Our use of the DGRBNuB in constraining the cosmic GRB rate is in the same spirit as efforts to constrain the cosmic star formation rate (and thereby the cosmic CCSN rate) by using the experimental upper limit (set by the Super-Kamiokande (SK) detector)~\\cite{SK_limit} on the predicted~\\cite{dsnub_original} low (few MeV) energy Diffuse Supernova Neutrino Background (DSNuB); see, for example, Refs.~\\cite{Ando-Sato_rev_04,Strigari_etal_05,Hopkins-Beacom_06}. Now that the cosmic SFR including its absolute normalization and thereby the cosmic CCSNe rate have got reasonably well determined by the recent high quality data from a variety of astronomical observations (see, e.g., ~\\cite{Hopkins-Beacom_06}) (which, by the way, predicts a DSNuB flux that is close to the SK upper limit, implying that the DSNuB is probably close to being detected in the near future), one can begin to think of using this SFR to constrain the ratio of the cosmic GRB rate to CCSNe rate by using the predicted DGRBNuB flux together with the recent upper limits on the diffuse high energy neutrino flux from neutrino telescopes. We should emphasize here that the upper limits derived in this paper actually refer to the ratio of the rate of GRBs to that of {\\it all} CCSNe including those of Type Ib/c and Type II, although SN-GRB associations observed so far involve SNe of Type Ib/c only. It is known, however, that Type II SNe probably constitute as much as $\\sim$ 75\\% of all CCSNe; see, e.g., \\cite{Cappellaro_etal_07}. Thus, one can get the constraint on the GRB-to-SNIb/c ratio from the GRB-to-CCSNe ratio we obtain here by multiplying the latter by a factor of $\\sim 4$. Conversely, for later comparison, we shall take the ``observed'' value of the ratio $\\fgrbccsn(0)$ to be in the range $2.5\\times (10^{-4}$ -- $10^{-3})$~\\cite{Guetta-DellaValle_06,Bissaldi_etal_07}. Below, we first briefly review the calculation of the DGRBNuB spectrum in section \\ref{sec:DGRBNuB_calc}. The resulting upper limits on $\\fgrbccsn$ obtained by comparing the DGRBNuB with the current upper limit from AMANDA-II experiment are discussed in section \\ref{sec:fgrbccsn_constraints} for various values of some of the relevant GRB parameters. Finally, in section \\ref{sec:summary} we summarize the main results and conclude. ", "conclusions": "} In this paper we have attempted to derive upper limits on the fraction $\\fgrbccsn$ of all stellar core-collapse events that give rise to GRBs, by using the current experimental upper limit on the high energy (TeV -- PeV) diffuse neutrino background given by the AMANDA-II experiment in the South Pole, under the assumption that GRBs are sources of such high energy neutrinos. High energy neutrinos are predicted to be produced within GRB jets through photopion production by protons and subsequent decay of the charged pions, provided protons are accelerated to ultrahigh energies at the internal shocks within GRB jets. In our calculation we have allowed for a possible evolution of the cosmic GRB rate relative to star formation rate. For a wide range of values of various parameters, the upper limits on $\\fgrbccsn(0)$ derived here from the AMANDA-II results are already more restrictive than the upper limit on this ratio inferred from other astronomical considerations, thus providing a useful independent probe of and constraint on the CCSN-GRB connection. The closeness of the upper limits on $\\fgrbccsn(0)$ derived here (in particular for the case of enhanced evolution of the GRB rate relative to the star formation rate at high redshifts) to the lower limit on this ratio inferred from various astronomical considerations seems to indicate that the predicted DGRBNuB flux should be detectable by the upcoming detectors such as IceCube which will have significantly improved sensitivity over that of AMANDA-II. On the other hand, non-detection of the DGRBNuB by the IceCube detector after three years of operation, for example, will give more stringent upper limits on $\\fgrbccsn$, but at the same time will also imply that either the values of $\\fgrbccsn$ inferred from direct astronomical observations have been significantly overestimated (which is possible, for example, due to incorrect estimates of the average GRB beaming factor) or that the assumption of proton acceleration to ultrahigh energies within GRB jets is invalid, or both of these. However, more precise determination of the distribution of some of the crucial GRB parameters such as the bulk flow Lorentz factor and variability timescale of the GRBs will be needed to reliably calculate the expected contribution of the GRBs to the high energy diffuse neutrino background, and thereby to determine the upper limits on $\\fgrbccsn$ more reliably. To conclude, then, the up-coming large volume neutrino telescopes hold immense promise of yielding significant information both on the nature of the fundamental physical process of particle acceleration in GRB sources as well as on the rate of occurrence of these events in the Universe. One of us (PB) wishes to thank Nayantara Gupta for helpful clarifications." }, "0710/0710.4029_arXiv.txt": { "abstract": "The cold dark matter (CDM) scenario generically predicts the existence of triaxial dark matter haloes which contain notable amounts of substructure. However, analytical halo models with smooth, spherically symmetric density profiles are routinely adopted in the modelling of light propagation effects through such objects. In this paper, we address the biases introduced by this procedure by comparing the surface mass densities of actual N-body haloes against the widely used analytical model suggested by Navarro, Frenk and White (1996) (NFW). We conduct our analysis in the redshift range of 0.0 - 1.5. In cluster sized haloes, we find that triaxiality can cause scatter in the surface mass density of the haloes up to $\\sigma_+ = +60 \\%$ and $\\sigma_- = -70 \\%$, where the 1-$\\sigma$ limits are relative to the analytical NFW model given value. Subhaloes can increase this scatter to $\\sigma_+ = +70 \\%$ and $\\sigma_- = -80 \\%$. In galaxy sized haloes, the triaxial scatter can be as high as $\\sigma_+ = +80 \\%$ and $\\sigma_- = -70 \\%$, and with subhaloes the values can change to $\\sigma_+ = +40 \\%$ and $\\sigma_- = -80 \\%$. We present an analytical model for the surface mass density scatter as a function of distance to the halo centre, halo redshift and halo mass. The analytical description enables one to investigate the reliability of results obtained with simplified halo models. Additionally, it provides the means to add simulated surface density scatter to analytical density profiles. As an example, we discuss the impact of our results on the calculation of microlensing optical depths for MACHOs in CDM haloes. ", "introduction": "The cold dark matter model, in which the non-baryonic part of the dark matter is assumed to consist of particles that were non-relativistic already at the time of decoupling, and that interact predominantly through gravity, has been very successful in explaining the formation of large-scale structures in the Universe \\citep[see e.g.][ for a review]{Primack}. In this scenario, both galaxies and galaxy clusters are hosted by CDM haloes, which formed hierarchically through mergers of smaller subunits. Even though N-body simulations generically predict CDM haloes to be triaxial \\citep[e.g.][]{Jing & Suto} with substantial amounts of substructures left over from the merging process \\citep[e.g.][]{Moore et al.}, simplified halo models are often adopted in the modelling of light propagation through such objects. The most common approach is to treat dark matter haloes as spherical objects with smooth density profiles, usually either of the NFW \\citep{NFW} form, some generalization thereof \\citep{Zhao}, or that of a cored or singular isothermal sphere. The light emitted from high-redshift objects such as quasars, supernovae, gamma-ray bursts, galaxies and galaxy clusters will typically have to pass through many dark matter haloes before reaching an observer on Earth. Several investigations have already indicated that smooth and/or spherical halo models may lead to incorrect results when treating the gravitational lensing effects associated with such foreground mass condensations \\citep[e.g][]{Bartelmann & Weiss,Dalal et al.,Oguri & Keeton,Hennawi et al.} More realistic features like triaxiality and substructures can be included in gravitational lens calculations either by employing N-body simulations directly \\citep[e.g.][]{Bartelmann & Weiss,Seljak & Holz,Holopainen et al.} or by using analytical expressions for the halo shapes \\citep[e.g.][]{Kochanek,Golse & Kneib,Evans & Hunter,Chae} and subhalo properties \\citep[e.g.][]{Oguri b,Zackrisson & Riehm}. While N-body simulations often represent the safest choice, the approach is computationally demanding and does not always allow one to identify the features of the mass distribution responsible for a specific lensing effect. Methods which bring simple, analytical halo models into contact with the full phenomenology of the N-body simulations are therefore highly desirable. In this paper, we focus on the projected mass density of CDM haloes as a function of distance from the halo centre. There are several situations in gravitational lensing when realistic estimates of the surface mass density (i.e. convergence) along a given line of sight through a dark halo may be important. Examples include the calculation of image separations in strong lensing by subhaloes located in the external potential of its host halo \\citep{Oguri b}, attempts to correct the luminosities of supernovae type Ia for the magnification by foreground haloes \\citep[e.g.][]{Gunnarsson} and estimates of the distribution of microlensing optical depths for high-redshift MACHOs \\citep[e.g.][]{Wyithe & Turner,Zackrisson & Riehm}. Other applications include the assessments of light propagation effects in models with non-zero coupling betweeen dark matter particles and photons \\citep[e.g.][]{Profumo & Sigurdson}. Here, we use high-resolution, dissipationless N-body simulations of CDM haloes to investigate the errors in surface mass density introduced by treating these objects as spherical with smooth density profiles of the NFW type. Simple relations for the surface mass density error as a function of halo redshift and distance to the halo centre are presented, making it easy to investigate the reliablitiy of results obtained with simplified halo models. On a related note, \\citet{KnebeWiessner} recently investigated the error introduced by spherically averaging an elliptical mass distribution. They found that for axis ratios typical for cosmological dark matter haloes, the variance in the local density can be as large as 50\\% in the outer parts. The current paper examines the problem of halo triaxiality from a slightly different point of view. The N-body simulations used are described in Section 2. In Section 3, we describe the methods for extracting the halo sample. In Section 4, we compare the CDM surface mass densities obtained along random sightlines through the N-body haloes to the corresponding results obtained from smooth and spherical NFW models fitted to the same haloes. Section 5 presents a set of simple relations for the surface mass density errors introduced by this procedure as a function of distance to the halo centre and halo redshift. Section 6 discusses how these relations may be used in the context of optical depth estimates for MACHO microlensing. A number of caveats are discussed in Section 7. Section 8 summarizes our findings. ", "conclusions": "The results presented here do suffer from a number of shortcomings which should be pointed out. The simulations used are dissipationless. In reality, dark matter haloes contain baryons, and the dissipation and feedback associated with these will inevitably affect the overall potential of the system, and thereby the spatial distribution of the CDM. According to current models, baryonic cooling will increase the central density of the CDM \\citep[e.g.][]{Gnedin et al.} and also make the halo more spherical \\citep{Kazantzidis et al.}. The significance of these effects are, however, still difficult to predict reliably, as the gas dynamical simulations involved still suffer from so-called ``overmerging'' problems \\citep[e.g.][]{Balogh et al., Springel & Hernquist}. When calculating the surface mass density profiles, we have moreover considered only the matter present within $r_\\mathrm{vir}$ of each halo, whereas simulations have shown that galaxy sized CDM haloes extend at least out to 2--3$r_\\mathrm{vir}$ \\citep{Prada et al.}. We restricted our analysis to $r_\\mathrm{vir}$ because it becomes increasingly demanding to separate halo particles from the background the further one wants to extend the analysis. Even in the presented case, we need to separate particles out to $\\sim 1.5 r_\\mathrm{vir}$ because the smoothed particles extend their influence inside the virial radius region even though they are positioned outside it. We tested our method out to 3$r_\\mathrm{vir}$, but the number counts of the halo particles at those distances are too low to produce reliable results. Our halo sample does not have the resolution needed for extending the analysis further than $r_\\mathrm{vir}$ safely. The most significant limitation of this paper is the small number of haloes in our analysis. This is of course due to the limited resolution of the cosmological simulations we had access to. We would like to repeat our analysis with a more complete statistical sample of haloes, which would hopefully confirm our analytical description with smaller error bars. We also note that our analytical description of the surface mass density is more reliable in the case in which subhaloes are {\\it excluded}. This is because subhaloes can introduce significant mass peaks to some radial bins. These peaks can lead to unwanted effects in the log-normal fitting procedure which is designed to handle relatively smooth and continuous mass distributions within a bin. Large subhalos can also disturb the NFW fits, at last in the low density regions. Thus, the use of the models which include subhaloes is discouraged." }, "0710/0710.1110_arXiv.txt": { "abstract": "We examine the effects Lorentz violation on observations of cosmic microwave background radiation. In particular, we focus on changes in polarization caused by vacuum birefringence. We place stringent constraints on previously untested violations. ", "introduction": " ", "conclusions": "" }, "0710/0710.3862_arXiv.txt": { "abstract": "We study a large-scale instability in a sheared nonhelical turbulence that causes generation of large-scale vorticity. Three types of the background large-scale flows are considered, i.e., the Couette and Poiseuille flows in a small-scale homogeneous turbulence, and the \"log-linear\" velocity shear in an inhomogeneous turbulence. It is known that laminar plane Couette flow and antisymmetric mode of laminar plane Poiseuille flow are stable with respect to small perturbations for any Reynolds numbers. We demonstrate that in a small-scale turbulence under certain conditions the large-scale Couette and Poiseuille flows are unstable due to the large-scale instability. This instability causes formation of large-scale vortical structures stretched along the mean sheared velocity. The growth rate of the large-scale instability for the \"log-linear\" velocity shear is much larger than that for the Couette and Poiseuille background flows. We have found a turbulent analogue of the Tollmien-Schlichting waves in a small-scale sheared turbulence. A mechanism of excitation of turbulent Tollmien-Schlichting waves is associated with a combined effect of the turbulent Reynolds stress-induced generation of perturbations of the mean vorticity and the background sheared motions. These waves can be excited even in a plane Couette flow imposed on a small-scale turbulence when perturbations of mean velocity depend on three spatial coordinates. The energy of these waves is supplied by the small-scale sheared turbulence. ", "introduction": "Large-scale vortical structures are universal features observed in geophysical, astrophysical and laboratory flows (see, e.g., \\cite{L83,P87,C94,GLM97,T98,RAO98}). Formation of vortical structures is related to the Prandtl secondary flows (see, e.g., \\cite{P52,T56,P70,B87}). A lateral stretching (or ''skewing\") by an existing shear generates streamwise vorticity that results in formation of the first kind of the Prandtl secondary flows. In turbulent flow the large-scale vorticity is generated by the divergence of the Reynolds stresses. This mechanism determines the second kind of the Prandtl turbulent secondary flows \\cite{B87}. The generation of large-scale vorticity in a homogeneous nonhelical turbulence with an imposed large-scale linear velocity shear has been recently studied in \\cite{EKR03}. Let us discuss a mechanism of this phenomenon. The equation for the mean vorticity ${\\bf W} = \\bec{\\nabla} {\\bf \\times} {\\bf U}$ read \\begin{eqnarray} {\\partial {\\bf W} \\over \\partial t} = \\bec{\\nabla} {\\bf \\times} ({\\bf U} {\\bf \\times} {\\bf W} + {\\bf F} - \\nu \\bec{\\nabla} {\\bf \\times} {\\bf W}) \\;, \\label{W10} \\end{eqnarray} where ${\\bf U}$ is the mean fluid velocity, ${\\bf F}_i = - \\nabla_j \\, \\langle u_i u_j \\rangle$ is the effective force caused by velocity fluctuations, ${\\bf u}$, and $ \\nu$ is the kinematic viscosity. The first term, ${\\bf U} {\\bf \\times} {\\bf W}$, in Eq.~(\\ref{W10}) determines laminar effects of the mean vorticity production caused by the sheared motions, while the effective force ${\\bf F}$ determines the turbulent effects on the mean fluid flow. Let us consider a simple large-scale linear velocity shear ${\\bf U}^{(s)} = (0, Sx, 0)$ imposed on the small-scale nonhelical turbulence. The equation for the perturbations of the mean vorticity, $\\tilde{\\bf W} = (\\tilde{W}_x(z), \\tilde{W}_y(z), 0)$, reads \\begin{eqnarray} {\\partial \\tilde{W}_x \\over \\partial t} &=& S \\, \\tilde{W}_y + \\nu_{_{T}} \\tilde{W}''_x \\;, \\label{E2}\\\\ {\\partial \\tilde{W}_y \\over \\partial t} &=& - \\beta_0 \\, S \\, l_0^2 \\, \\tilde{W}''_x + \\nu_{_{T}} \\tilde{W}''_y \\;, \\label{E3} \\end{eqnarray} (see \\cite{EKR03}), where $\\tilde{W}'' = \\partial^2 \\tilde{W} /\\partial z^2$, $\\, \\nu_{_{T}}$ is the turbulent viscosity, $l_0$ is the maximum scale of turbulent motions and the parameter $\\beta_0$ is of the order of 1, and depends on the scaling exponent of the correlation time of the turbulent velocity field (see Sect. II). A solution of Eqs.~(\\ref{E2}) and~(\\ref{E3}) has the form $ \\propto \\exp(\\gamma t + i K_z z)$, where the growth rate of the large-scale instability is given by $\\gamma = \\sqrt{\\beta_0} \\, S \\, l_0 \\, K_z - \\nu_{_{T}} \\, K_z^2$ and $K_z$ is the wave number. The maximum growth rate of perturbations of the mean vorticity, $ \\gamma_{\\rm max} = \\beta_0 \\, (S \\, l_0)^2 / 4 \\nu_{_{T}}$, is attained at $ K_z = K_m = \\sqrt{\\beta_0} \\, S \\, l_0 /2 \\nu_{_{T}}$. This corresponds to the ratio $\\tilde{W}_y / \\tilde{W}_x = \\sqrt{\\beta_0} \\, l_0 \\, K_m \\approx S \\, \\tau_0$, where the time $\\tau_0 = l_{0} / u_0$ and $u_0$ is the characteristic turbulent velocity in the maximum scale $l_{0}$ of turbulent motions. Note that in a laminar flow this instability does not occur. The mechanism of this instability is as follows (see \\cite{EKR03} for details). The first term, $S \\tilde{W}_y = ({\\bf W}^{(s)}\\cdot\\bec{\\nabla})~\\tilde{U}_x$, in Eq.~(\\ref{E2}) determines a ''skew-induced\" generation of perturbations of the mean vorticity $\\tilde{W}_x$ by stretching of the equilibrium mean vorticity ${\\bf W}^{(s)}= (0,0,S)$, where $\\tilde{\\bf U}$ are the perturbations of the mean velocity. In particular, the mean vorticity $\\tilde{W}_x {\\bf e}_x$ is generated from $\\tilde{W}_y {\\bf e}_y$ by equilibrium shear motions with the mean vorticity ${\\bf W}^{(s)}$, whereby $\\tilde{W}_x {\\bf e}_x \\propto ({\\bf W}^{(s)} \\cdot \\bec{\\nabla}) \\tilde{U}_x {\\bf e}_x \\propto \\tilde{W}_y {\\bf e}_y \\times {\\bf W}^{(s)} $. Here ${\\bf e}_x$, ${\\bf e}_y$ and ${\\bf e}_z$ are the unit vectors along $x$, $y$ and $z$ axes, respectively. On the other hand, the first term, $- \\beta_0 \\, S \\, l_0^2 \\, \\tilde{W}''_x$, in Eq.~(\\ref{E3}) determines a ''Reynolds stress-induced\" generation of perturbations of the mean vorticity $\\tilde{W}_y$ by the Reynolds stresses. In particular, this term is determined by $ (\\bec{\\nabla} {\\bf \\times} {\\bf F})_y$. This implies that the component of the mean vorticity $\\tilde{W}_y {\\bf e}_y $ is generated by an effective anisotropic viscous term $ \\propto - l_0^2 \\, \\Delta \\, (\\tilde{W}_x {\\bf e}_x \\cdot \\bec{\\nabla}) \\, {U}^{(s)}(x) {\\bf e}_y \\propto - l_0^2 \\, S \\, \\tilde{W}''_x {\\bf e}_y .$ This instability is caused by a combined effect of the sheared motions (''skew-induced\" generation) and the ''Reynolds stress-induced\" generation of perturbations of the mean vorticity. The mechanism for this large-scale instability in a sheared nonhelical homogeneous turbulence is different from that discussed in \\cite{MST83,KMT91,CMP94}, where the generation of large-scale vorticity in the helical turbulence occurs due to hydrodynamic alpha effect. The latter effect is associated with the hydrodynamic helicity of turbulent flow. In a nonhelical homogeneous turbulence this effect does not occur. The large-scale instability in a nonhelical homogeneous turbulence has been studied in \\cite{EKR03} only for a simple case of unbounded turbulence with an imposed linear velocity shear and when the perturbations of the mean vorticity depend on one spatial variable $z$. In this study the theoretical approach proposed in \\cite{EKR03} is further developed and applied for comprehensive investigation of the large-scale instability for different situations with nonuniform shear, inhomogeneous turbulence and a more general form of the perturbations of the mean vorticity $\\tilde{\\bf W}({\\bf r})$ that depends on three spatial variables. In the present study we consider three types of the background large-scale flows, i.e., the Couette flow (linear velocity shear) and Poiseuille flow (quadratic velocity shear) in a small-scale homogeneous turbulence, and the \"log-linear\" velocity shear in an inhomogeneous turbulence. We have derived new mean-field equations for perturbations of large-scale velocity which depend on three spatial coordinates in a small-scale sheared turbulence, for a nonuniform background large-scale velocity shear and for an arbitrary scaling of the correlation time $\\tau(k)$ of the turbulent velocity field. The stability of the laminar Couette and Poiseuille flows in a problem of transition to turbulence has been studied in a number of publications (see, e.g., \\cite{DR81,SH01,CJJ03,BOH88,REM03,ESH07}, and references therein). It is known that laminar plane Couette flow and antisymmetric mode of laminar plane Poiseuille flow are stable with respect to small perturbations for any Reynolds numbers. A symmetric mode of laminar plane Poiseuille flow is stable when the Reynolds number is less than 5772 \\cite{CJJ03}. In laminar flows the Tollmien-Schlichting waves can be excited. The molecular viscosity plays a destabilizing role in laminar flows which promotes the excitation of the Tollmien-Schlichting waves (see, e.g., \\cite{SH01}). These waves are growing solutions of the Orr-Sommerfeld equation. In the present study we have found a turbulent analogue of the Tollmien-Schlichting waves. These waves are excited by a small-scale sheared turbulence, i.e., by a combined effect of the turbulent Reynolds stress-induced generation of perturbations of the mean vorticity and the background sheared motions. The energy of these waves is supplied by the small-scale sheared turbulence. We demonstrate that the off-diagonal terms in the turbulent viscosity tensor play a crucial role in the excitation of the turbulent Tollmien-Schlichting waves. These waves can be excited even in a plane Couette flow imposed on a small-scale turbulence when perturbations of velocity depend on three spatial coordinates. When perturbations of large-scale velocity depend on one or two spatial coordinates the turbulent Tollmien-Schlichting waves can not be excited in a sheared turbulence. In the present study we show that the large-scale Couette and Poiseuille flows imposed on a small-scale turbulence can be unstable with respect to small perturbations. The critical effective Reynolds number (based on turbulent viscosity) required for the excitation of this large-scale instability, is of the order of 200. This paper is organized as follows. In Sect. II the governing equations are formulated. In Sect. III we consider a homogeneous turbulence with a large-scale linear velocity shear (Couette flow), while in Sect. IV we study a homogeneous turbulence with a large-scale quadratic velocity shear (Poiseuille flow). In Sect. V we investigate formation of large-scale vortical structures in an inhomogeneous turbulence with an imposed nonuniform velocity shear. Finally, we draw conclusions in Sec.~VI. ", "conclusions": "In this study the theoretical approach proposed in \\cite{EKR03} is further developed and applied to investigate the large-scale instability in a nonhelical turbulence with a nonuniform shear and a more general form of the perturbations of the mean vorticity. In particular, we consider three types of the background large-scale sheared flows imposed on small-scale turbulence: Couette flow (linear velocity shear) and Poiseuille flow (quadratic velocity shear) in a small-scale homogeneous turbulence, and a more complicated nonuniform velocity shear with the logarithmic velocity profile near the boundaries matched with the linear shear velocity for the central part of the background flow. This nonuniform velocity shear is imposed on an inhomogeneous turbulence. The latter flow is typical for the atmospheric boundary layer. We show that the large-scale Couette and Poiseuille flows imposed on a small-scale turbulence are unstable with respect to small perturbations due to the excitation of the large-scale instability. This instability causes generation of large-scale vorticity and formation of large-scale vortical structures. The size of the formed vortical structures in the direction of the background velocity shear is much larger than the sizes of the structures in the directions perpendicular to the velocity shear. Therefore, the large-scale structures formed during this instability are stretched along the mean sheared velocity. Increase of shear promotes the large-scale instability. The thresholds for the excitation of the large-scale instability in the value of shear and the aspect ratio of structures for Poiseuille background flow are larger than that for the Couette background flow. The growth rate of the large-scale instability for the inhomogeneous turbulence with the \"log-linear\" velocity shear is much larger than that for the Couette and Poiseuille background flows. The characteristic spatial and time scales for the instability are much larger than the characteristic turbulent scales. This justifies separation of scales which is required for the validity of the mean-field theory applied in the present study. The large-scale instability results in excitation of the turbulent Tollmien-Schlichting waves. The mechanism for the excitation of these waves is different from that for the Tollmien-Schlichting waves in laminar flows. In particular, the molecular viscosity plays a crucial role in the excitation of the Tollmien-Schlichting waves in laminar flows. Contrary, the turbulent Tollmien-Schlichting waves are excited by a combined effect of the turbulent Reynolds stress-induced generation of perturbations of the mean vorticity and the background sheared motions. The energy of these waves is supplied by the small-scale sheared turbulence, and the off-diagonal terms in the turbulent viscosity tensor play a crucial role in the excitation of the turbulent Tollmien-Schlichting waves. Note that this study is principally different from the problems of transition to turbulence whereby the stability of the laminar Couette and Poiseuille flows are investigated (see, e.g., \\cite{DR81,SH01,CJJ03,BOH88,REM03,ESH07}, and references therein). Here we do not analyze a transition to turbulence. We study the large-scale instability caused by an effect of the small-scale anisotropic turbulence on the mean flow. This anisotropic turbulence is produced by an interaction of equilibrium large-scale Couette or Poiseuille flows with a small-scale isotropic background turbulence produced by, e.g., a steering force. The anisotropic velocity fluctuations are generated by tangling of the mean-velocity gradients with the velocity fluctuations of the background turbulence \\cite{EKR03,EKRZ02}. The \"tangling\" mechanism is an universal phenomenon that was introduced in \\cite{W57,BH59} for a passive scalar and in \\cite{G60,M61} for a passive vector (magnetic field). The Reynolds stresses in a turbulent flow with a mean velocity shear is another example of tangling anisotropic fluctuations \\cite{L67}. For instance, these velocity fluctuations are anisotropic in the presence of shear and have a steeper spectrum $\\propto k^{-7/3}$ than, e.g., a Kolmogorov background turbulence (see, e.g., \\cite{L67,WC72,SV94,IY02,EKRZ02}). The anisotropic velocity fluctuations determine the effective force and the Reynolds stresses in Eq.~(\\ref{B15}). This is the reason for the new terms $\\propto \\beta_n \\, l_0^2$ appearing in Eqs.~(\\ref{AA1})-(\\ref{A2}). The obtained results in this study may be of relevance in different turbulent astrophysical, geophysical and industrial flows. Turbulence with a large-scale velocity shear is a universal feature in astrophysics and geophysics. In particular, the analyzed effects may be important, e.g., in accretion disks, extragalactic clusters, merged protostellar and protogalactic clouds. Sheared motions between interacting clouds can cause an excitation of the large-scale instability which results in generation of the mean vorticity and formation of large-scale vortical structures (see, e.g., \\cite{P80,ZN83,C93}). Dust particles can be trapped by the vortical structures to enhance agglomeration of material and formation of particle clusters \\cite{BS95,BR98,EKR98,CH00,JAB04}. The suggested mechanism can be used in the analysis of the flows associated with Prandtl's turbulent secondary flows (see, e.g., \\cite{P52,B87}). However, in this study we have investigated only simple physical mechanisms to describe an initial (linear) stage of the formation of vortical structures. The simple models considered in this study can only mimic the flows associated with turbulent secondary flows. Clearly, the comprehensive numerical simulations of the nonlinear problem are required for quantitative description of the turbulent secondary flows." }, "0710/0710.2864_arXiv.txt": { "abstract": "We present results from two high--contrast imaging surveys that exploit a novel technique, L--band angular differential imaging. Our first survey targeted 21 young stars in the $\\beta$~Pic and Tuc--Hor moving groups with VLT/NACO reaching typical sensitivities of $<$1~M$_{\\mathrm Jup}$ at $r>20$~AU. The statistical analysis of the null result demonstrates that the giant planet population is truncated at 30~AU or less (90\\% confidence level). Our second, on--going MMT/Clio survey utilizes the unique sensitivity achieved in the L--band for old planets to probe all M--dwarf stars within 6~pc. The proximity of these targets enables direct imaging of planets in orbits like Jupiter for the first time --- a key step for directly imaging giant planets. ", "introduction": "The study of exoplanetary systems is arguably the most rapidly developing field in modern astrophysics. Surprisingly, much progress has been made without directly imaging a single planet: radial velocity/microlensing and primary and secondary planet eclipses provide limited, but valuable insights. Direct imaging of planetary systems will have a fundamental impact on the field --- a single, low--resolution 3--5 $\\mu$m spectrum of a planet may carry more information than all existing Spitzer transit photometry combined. As was the case in the search for the first brown dwarf, or for radial velocity and planet transit techniques, achieving the first firm detection is a very difficult and often frustrating challenge. But these investments paid off rapidly by opening whole new classes of objects for study. ", "conclusions": "We present results from a novel high--contrast imaging technique. Our NACO survey of 21 nearby young stars demonstrates that the giant planet population does not extend beyond 30 AU and suggests a cut-off at radius $<$ 15 AU. Most previous imaging surveys have not detected planets because they targeted young stars ($>$ 20 pc) forcing them to probe orbital radii $>$20~AU. Our ongoing MMT 6pc volume--limited survey is probing --- for the first time --- the massive giant planet population around the closest stars with orbital radii $>$3." }, "0710/0710.5169_arXiv.txt": { "abstract": "We construct new models of black hole-neutron star binaries in quasiequilibrium circular orbits by solving Einstein's constraint equations in the conformal thin-sandwich decomposition together with the relativistic equations of hydrostationary equilibrium. We adopt maximal slicing, assume spatial conformal flatness, and impose equilibrium boundary conditions on an excision surface (i.e., the apparent horizon) to model the black hole. In our previous treatment we adopted a ``leading-order\" approximation for a parameter related to the black-hole spin in these boundary conditions to construct approximately nonspinning black holes. Here we improve on the models by computing the black hole's quasilocal spin angular momentum and setting it to zero. As before, we adopt a polytropic equation of state with adiabatic index $\\Gamma=2$ and assume the neutron star to be irrotational. In addition to recomputing several sequences for comparison with our earlier results, we study a wider range of neutron star masses and binary mass ratios. To locate the innermost stable circular orbit we search for turning points along both the binding energy and total angular momentum curves for these sequences. Unlike for our previous approximate boundary condition, these two minima now coincide. We also identify the formation of cusps on the neutron star surface, indicating the onset of tidal disruption. Comparing these two critical binary separations for different mass ratios and neutron star compactions we distinguish those regions that will lead to a tidal disruption of the neutron star from those that will result in the plunge into the black hole of a neutron star more or less intact, albeit distorted by tidal forces. ", "introduction": "Coalescing black hole-neutron star (BHNS) binaries, as well as other compact binaries composed of neutron stars and/or black holes, are among the most promising sources of gravitational waves for both ground-based \\cite{LIGO,GEO,TAMA,VIRGO} and space-based laser interferometers \\cite{LISA,DECIGO}. BHNS binary mergers are also candidate central engines of short-hard gamma-ray bursts (SGRBs) (see, e.g. \\cite{LeeR07} and references cited therein). The remnants of both BHNS binary mergers \\cite{FaberBST06,ShibaU0607} and binary neutron star mergers \\cite{ShibaT06,PriceR06,OechsJ06,HMNS} are feasible progenitors for SGRBs because both may result in black holes surrounded by hot, massive accretion disks with very little, if any, baryon contamination along the polar symmetry axis. Motivated by these factors, considerable effort has gone into the study of BHNS binaries. Most approaches to date assume Newtonian gravity in either some or all aspects of the calculation (see, e.g.~\\cite{Chand69,Fishb73,LaiRS93,LaiW96,TanigN96,Shiba96,UryuE99,WiggiL00,IshiiSM05,Mille05} for quasiequilibrium calculations and \\cite{Mashh75,CarteL83,Marck83,LeeK99,Lee00,RosswSW04,KobayLPM04,RantsKLR07} for dynamical simulations). More recently, several groups have also studied BHNS binaries in a fully relativistic framework, both for quasiequilibrium models \\cite{Mille01,BaumgSS04,TanigBFS05,TanigBFS06,Grand06,TanigBFS07} and dynamical simulations \\cite{FaberBSTR06,FaberBST06,SopueSL06,LofflRA06,ShibaU0607}. Our group has pursued a systematic approach to developing increasingly realistic models of BHNS binaries in quasiequilibrium circular orbits. Our first studies \\cite{BaumgSS04,TanigBFS05,FaberBSTR06,FaberBST06} assumed extreme mass ratios, i.e., black hole masses that are much greater than the neutron star mass. While this is a very natural first step from a computational point of view, binaries with comparable masses are much more interesting from the perspective of ground-based gravitational wave observations and for the launching of SGRBs. More recently we have therefore relaxed this assumption and have extended our results to the case of comparable-mass BHNS binaries \\cite{TanigBFS06,TanigBFS07}. Specifically, in \\cite{TanigBFS07} (hereafter Paper I) we constructed quasiequilibrium models by solving Einstein's constraint equations in the conformal thin-sandwich formalism, assuming conformal flatness and maximal slicing, together with the relativistic equations of hydrostationary equilibrium. We accounted for the black hole by excising a coordinate sphere and imposing the equilibrium black-hole boundary conditions of Cook and Pfeiffer \\cite{CookP04}. This original version implemented a ``leading-order\" approximation to nonspinning black holes, which equates an otherwise undetermined spin parameter $\\Omega_r$ that appears in the boundary condition for the shift vector with the orbital angular velocity seen by an inertial observer at infinity, $\\Omega$. As for the original irrotational binary black hole models of \\cite{CookP04}, this condition does not lead to simultaneous turning points of the binding energy and the total angular momentum in constant-mass sequences in Paper I. Such simultaneous turning points are expected for those sequences if they are truly in quasiequilibrium \\cite{dMOmegadJ}. An improvement over this condition, namely to iterate over $\\Omega_r$ until the quasilocal spin angular momentum of the black hole vanishes, was suggested and implemented for binary black holes by \\cite{CaudiCGP06}. In this paper we reconstruct quasiequilibrium models of BHNS binaries using the same techniques as in Paper I, but with the improved black hole spin angular velocity condition as suggested by \\cite{CaudiCGP06}. We then compute sequences of BHNS binaries in quasicircular orbits for a wider range of neutron star masses and binary mass ratios than in Paper I, focusing our attention on irrotational neutron stars orbiting nonspinning black holes. Here we focus only on the irrotational state for the neutron star because it is astrophysically considered to be more realistic in a BHNS binary \\cite{Kocha92,BildsC92,FaberBSTR06}. On the other hand, we will compute the case of spinning black holes in future work. As was the case for the irrotational black hole binaries constructed in \\cite{CaudiCGP06}, we find that this improved condition for the spin parameter of the black hole $\\Omega_r$ does lead to simultaneous turning points in the binding energy and the total angular momentum along constant-mass sequences. The paper is organized as follows. We briefly review the basic equations in Section II. We present numerical results in Section III, and outline some qualitative considerations concerning the fate of BHNS binaries in Section IV. In Section V we summarize our findings. Throughout this paper we adopt geometrized units with $G=c=1$, where $G$ denotes the gravitational constant and $c$ the speed of light. Latin and Greek indices denote purely spatial and spacetime components, respectively. ", "conclusions": "We have constructed new quasiequilibrium configurations of black hole-neutron star binaries in general relativity. We have solved the Einstein constraint equations in the conformal thin-sandwich formalism coupled with the equations of relativistic hydrostationary equilibrium. In Paper I, we set the spin angular velocity parameter of the black hole equal to that of the orbital angular velocity in order to produce a nonspinning black hole in the ``leading-order\" approximation \\cite{CookP04}, while in this paper we compute this parameter by requiring the quasilocal spin angular momentum of the black hole to be zero \\cite{CaudiCGP06}. We have also improved the formulation of the gravitational field equations and obtained more accurate results than in Paper I. As an indication of the improvements in these calculations, a post-Newtonian analysis predicts smaller binary eccentricities for these new BHNS models than for those computed in Paper I (\\cite{Will07}, compare \\cite{BertiIW07}). In \\cite{BertiIW07}, Berti {\\it et al.} fit numerical results for the binding energy and angular momentum of binaries in circular orbit to post-Newtonian expressions for binaries that are not necessarily in circular orbit. Deviations between the the two approaches then lead to non-zero eccentricities in the post-Newtonian expressions. These eccentricities are smaller for our new results than for those of Paper I. We also remark on another finding of \\cite{BertiIW07}, namely that for a given neutron star mass $\\bar{M}_{\\rm ADM,0}^{\\rm NS}$ and a given value of $\\Omega M_0$, the eccentricities in BHNS models, though small, are found to be larger than in binary neutron star models \\cite{TanigG0203}. This suggests a larger deviation from quasiequilibrium for BHNS binaries than binary neutron stars. But, for BHNS binaries with a mass ratio of $\\hat{q} = 5$, these parameters correspond to a larger binary separation than for binary neutron stars with a mass ratio of $\\hat{q} = 1$ (compare Eq.~(\\ref{eq:Kepler})). For similar numerical resources, this larger binary separation leads to a larger numerical error, which may explain the larger eccentricity found by \\cite{BertiIW07}, at least in part. In addition to recomputing several sequences we presented in Paper I, we have constructed sequences for a wider range of neutron star masses and binary mass ratios, employing a $\\Gamma=2$ polytropic neutron-star equation of state throughout. We computed several constant-mass sequences, for various mass ratios and neutron star compactions, and searched for the appearance of a cusp at the neutron star surface -- indicating the onset of tidal disruption -- and turning points on the binding energy and angular momentum curves -- identifying the ISCO. We also included some qualitative fits that allow for a simple prediction of those binary parameters separating these two different outcomes of binary coalescence. Unlike in our earlier findings, we found simultaneous turning points along the binding energy and angular momentum quasiequilibrium curves." }, "0710/0710.0785_arXiv.txt": { "abstract": "{\\it NeXT} (New X-ray Telescope) is the next Japanese X-ray astronomical satellite mission after the {\\it Suzaku} satellite. {\\it NeXT} aims to perform wide band imaging spectroscopy. Due to the successful development of a multilayer coated mirror, called a supermirror, {\\it NeXT} can focus X-rays in the energy range from 0.1~keV up to 80~keV. To cover this wide energy range, we are in the process of developing a hybrid X-ray camera, Wideband X-ray Imager (WXI) as a focal plane detector of the supermirror. The WXI consists of X-ray CCDs (SXI) and CdTe pixelized detectors (HXI), which cover the lower and higher X-ray energy bands of 0.1--80~keV, respectively. The X-ray CCDs of the SXI are stacked above the CdTe pixelized detectors of the HXI. The X-ray CCDs of the SXI detect soft X-rays below $\\sim 10$~keV and allow hard X-rays pass into the CdTe detectors of the HXI without loss. Thus, we have been developing a ``back-supportless CCD'' with a thick depletion layer, a thinned silicon wafer, and a back-supportless structure. In this paper, we report the development and performances of an evaluation model of CCD for the SXI, ``CCD-NeXT1''. We successfully fabricated two types of CCD-NeXT1, unthinned CCDs with 625-$\\mu$m thick wafer and 150-$\\mu \\rm m$ thick thinned CCDs. By omitting the polishing process when making the thinned CCDs, we confirmed that the polishing process does not impact the X-ray performance. In addition, we did not find significant differences in the X-ray performance between the two types of CCDs. The energy resolution and readout noise are $\\sim 140$~eV (FWHM) at 5.9~keV and $\\sim 5$ electrons (RMS), respectively. The estimated thickness of the depletion layer is $\\sim 80 ~\\mu \\rm m$. The performances almost satisfy the requirements of the baseline plan of the SXI. ", "introduction": "{\\it NeXT} (New X-ray Telescope) is the sixth Japanese X-ray astronomical satellite mission, which is proposed to be launched around 2012. A Hard X-ray Telescope (HXT), supermirror onboard {\\it NeXT} (see Tawara {\\it et~al.} (2003) \\cite{tawara03} and references therein), has a large collecting area for X-rays in the energy from 0.1 to 80~keV. In particular, the HXT has a high reflectivity even in the hard X-ray band above 10~keV. Previous satellite missions have not had X-ray focusing optics capable of observations in this band. {\\it NeXT} is designed to be the first to perform imaging and spectroscopic observations in the energy band above 10~keV. In order to meet the energy range covered by the HXT, we have been developing a Wideband X-ray Imager (WXI). The first successful space flight use of X-ray CCDs as photon counting and spectroscopic imagers was the SIS aboard $ASCA$ \\cite{byrke91}. Since then, X-ray CCDs have become standard focal plane detectors for X-ray telescopes in the X-ray energy band of 0.1--10~keV, and have been adopted as the principal detectors of recent X-ray observatories such as the ACIS of {\\it Chandra} \\cite{garmire03}, the EPIC of {\\it XMM-Newton} \\cite{struder01,turner01}, and the XIS of {\\it Suzaku} \\cite{koyama07} because X-ray CCDs have well balanced spectroscopic, imaging, and time resolution performances. However, to achieve a quantum efficiency of 10\\% for X-rays with an energy of 40~keV, a $\\sim 1000~{\\rm \\mu m}$ depletion layer is required, which is nearly impossible. A detector with a high-Z material is essential for observations in the hard X-ray band above 10--20~keV. On the other hand, the performances of imaging and spectroscopy below 10~keV of high-Z solid detectors such as CdTe detectors are poorer than those of X-ray CCDs, suggesting that a single detector cannot cover the entire 0.1--80~keV band with the best X-ray performance. Thus, we have been developing a hybrid camera, the Wide band X-ray Imager (WXI), which combines an X-ray CCD and a CdTe pixelized detector \\cite{Takahashi1999,tsuru01,tsuru04,takahashi04}. Holland (2003) \\cite{adholland03} has also reported the first laboratory demonstration of such a hybrid detector with a thinned X-ray CCD, which is operated in front of CZT detector. The WXI consists of two sub-instruments; the Soft X-ray Imager (SXI) and the Hard X-ray Imager (HXI); overviews can be found in Tsuru {\\it et~al.} (2005) \\cite{tsuru05} and Takahashi {\\it et~al.} (2004) \\cite{takahashi04s}, respectively. The SXI and HXI are the upper and lower parts of the WXI, respectively. The SXI consists of X-ray CCDs with a thick depletion layer for the lower energy band below 10--20~keV. The HXI is based on CdTe pixelized detectors, which cover the hard X-rays above 10--20~keV. We have developed a new type of CCD for the SXI, a ``Back-Supportless CCD'' (BS-CCD), in which the back supporting package under the imaging area of the CCD is removed. Most X-rays absorbed in the field-free region of the CCD are undetected and lost. Hence, we also removed the field-free region as much as possible. The BS-CCD of the SXI is placed over the CdTe pixelized detectors of the HXI. Soft X-rays are detected in the BS-CCD, while hard X-rays penetrate through the BS-CCD and are detected by the CdTe pixelized detectors. Thus, both soft and hard X-rays are detected without loss. As previously reported by Tsuru {\\it et~al.} (2005) \\cite{tsuru05} in detail, we have been developing a BS-CCD for the SXI following two plans of a rather conservative ``baseline plan'' and an innovative ``goal plan'' in parallel. Table~\\ref{tab:SXIgoala} shows the specifications of the BS-CCD of the two plans along with those of the XIS \\cite{koyama07}, which is one of the most excellent X-ray CCDs currently in orbit. In the goal plan, we realize a fully-depleted back-illuminated type of BS-CCD with a very thick depletion layer of $\\sim 200~{\\rm \\mu m}$ by adopting a p-channel device. We have already successfully fabricated test devices with a full depletion layer, which was $\\sim$ 200-$\\mu{\\rm m}$ thick. The details and status of the developments are reported elsewhere \\cite{kamata04,takagi05,kamata06,matuura06,takagi06,tsuru06}. In the baseline plan, we developed BS-CCDs based on a natural extension of our successful developments of CCD-CREST/CREST2 and MAXI-CCD in order to minimize the risk involved with their development \\cite{bamba01,tsunemi05,tomida00,miyata02}. We adopted a front-illuminated type of CCD mainly due to the manufacturing process \\cite{takagi05}. The thickness of the depletion layer is designed to be 70--80$~{\\rm \\mu m}$ or more. We have already successfully developed a small test model of BS-CCD and confirmed that the thinning processes of the wafer and the back-supportless structure do not degrade the performance \\cite{takagi05}. After the successfully developing the small test model, we constructed an evaluation model, ``CCD-NeXT1''. Following CCD-NeXT1, we will develop a flight model, ``CCD-NeXT2'', which matches the specifications of the SXI shown in Table~1. In this paper, we report the development and the performance of CCD-NeXT1\\footnote{Note that part of the results reported herein have already been reported as a contributing paper of a SPIE conference \\cite{OzawaSPIE06_CCD-NeXT1}.}. ", "conclusions": "\\begin{itemize} \\item Following the successful development of the test models of BS-CCD for SXI onboard the {\\it NeXT} satellite, we developed an evaluation model, CCD-NeXT1. We fabricated two types of CCD-NeXT1. One was a type of un-thinned CCD with 625-$\\mu$m thick wafers. The other type was a thinned CCD with 150-$\\mu \\rm m$ thickness. \\item We processed thinned CCD-NeXT1 devices by omitting the polishing process in the thinning process. The evaluation of the devices confirmed that omitting the polishing process did not impact on the X-ray performance. \\item We did not observe a significant difference in the X-ray performance of the unthinned and the thinned devices. The energy resolution and the readout noise were $\\sim$140~eV (FWHM) at 5.9~keV and $\\sim$5 electrons (RMS), respectively. The detection efficiency of the $\\rm ^{109}Cd$ photons indicates that the depletion layer is $\\sim$ 80-${\\rm \\mu m}$ thick. This performance meets the requirements for the baseline plan of the SXI. \\end{itemize}" }, "0710/0710.2780_arXiv.txt": { "abstract": "}[2]{{\\footnotesize\\begin{center}ABSTRACT\\end{center} \\vspace{1mm}\\par#1\\par \\noindent {~}{\\it #2}}} \\newcommand{\\TabCap}[2]{\\begin{center}\\parbox[t]{#1}{\\begin{center} \\small {\\spaceskip 2pt plus 1pt minus 1pt T a b l e} \\refstepcounter{table}\\thetable \\\\[2mm] \\footnotesize #2 \\end{center}}\\end{center}} \\newcommand{\\TableSep}[2]{\\begin{table}[p]\\vspace{#1} \\TabCap{#2}\\end{table}} \\newcommand{\\FigCap}[1]{\\footnotesize\\par\\noindent Fig.\\ % \\refstepcounter{figure}\\thefigure. #1\\par} \\newcommand{\\TableFont}{\\footnotesize} \\newcommand{\\TableFontIt}{\\ttit} \\newcommand{\\SetTableFont}[1]{\\renewcommand{\\TableFont}{#1}} \\newcommand{\\MakeTable}[4]{\\begin{table}[p]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newcommand{\\MakeTableTop}[4]{\\begin{table}[t]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newcommand{\\MakeTableSep}[4]{\\begin{table}[p]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newcommand{\\TabCapp}[2]{\\begin{center}\\parbox[t]{#1}{\\centerline{ \\small {\\spaceskip 2pt plus 1pt minus 1pt T a b l e} \\refstepcounter{table}\\thetable} \\vskip2mm \\centerline{\\footnotesize #2}} \\vskip3mm \\end{center}} \\newcommand{\\MakeTableSepp}[4]{\\begin{table}[p]\\TabCapp{#2}{#3}\\vspace*{-.7cm} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newfont{\\bb}{ptmbi8t at 12pt} \\newfont{\\bbb}{cmbxti10} \\newfont{\\bbbb}{cmbxti10 at 9pt} \\newcommand{\\uprule}{\\rule{0pt}{2.5ex}} \\newcommand{\\douprule}{\\rule[-2ex]{0pt}{4.5ex}} \\newcommand{\\dorule}{\\rule[-2ex]{0pt}{2ex}} \\def\\thefootnote{\\fnsymbol{footnote}} \\newenvironment{references}% { \\footnotesize \\frenchspacing \\renewcommand{\\thesection}{} \\renewcommand{\\in}{{\\rm in }} \\renewcommand{\\AA}{Astron.\\ Astrophys.} \\newcommand{\\AAS}{Astron.~Astrophys.~Suppl.~Ser.} \\newcommand{\\ApJ}{Astrophys.\\ J.} \\newcommand{\\ApJS}{Astrophys.\\ J.~Suppl.~Ser.} \\newcommand{\\ApJL}{Astrophys.\\ J.~Letters} \\newcommand{\\AJ}{Astron.\\ J.} \\newcommand{\\IBVS}{IBVS} \\newcommand{\\PASP}{P.A.S.P.} \\newcommand{\\Acta}{Acta Astron.} \\newcommand{\\MNRAS}{MNRAS} \\renewcommand{\\and}{{\\rm and }} {Period--luminosity (PL) relations of variable red giants in the Large (LMC) and Small Magellanic Clouds (SMC) are presented. The PL diagrams are plotted in three planes: $\\log P$--$K_S$, $\\log P$--$W_{JK}$, and $\\log P$--$W_I$, where $W_{JK}$ and $W_I$ are reddening free Wesenheit indices. Fourteen {\\it PL} sequences are distinguishable, and some of them consist of three closely spaced ridges. Each of the sequences is fitted with a linear or quadratic function. The similarities and differences between the {\\it PL} relations in both galaxies are discussed for four types of red giant variability: OGLE Small Amplitude Red Giants (OSARGs), Miras and Semiregular Variables (SRVs), Long Secondary Periods (LSPs) and ellipsoidal variables. We propose a new method of separating OSARGs from non-variable stars and SRVs. The method employs the position in the reddening-free {\\it PL} diagrams and the characteristic period ratios of these multiperiodic variables. The {\\it PL} relations for the LMC OSARG are compared with the calculated relations for RGB models along isochrones of relevant ages and metallicities. We also compare measured periods and amplitudes of the OSARGs with predictions based on the relations valid for less luminous solar-like pulsators. Miras and SRVs seem to follow {\\it PL} relation of the same slopes in the LMC and SMC, while for LSP and ellipsoidal variables slopes in both galaxies are different. The {\\it PL} sequences defined by LSP variables and binary systems overlap in the whole range of analyzed wavebands. We put forward new arguments for the binary star scenario as an explanation of the LSP variability and elaborate on it further. The measured pulsation to orbital period ratio implies nearly constant ratio of the star radius to orbital distance, $R/A\\approx0.4$, as we find. Combined effect of tidal friction and mass loss enhanced by the low-mass companion may explain why such a value is preferred.}{Stars: AGB and post-AGB -- Stars: late-type -- Stars: oscillations -- Magellanic Clouds} ", "introduction": "The first attempt to determine period--luminosity (PL) relation for long period variables (LPVs) was made by Gerasimovi\\v{c} (1928), who noticed that Mira stars with longer periods are on average fainter at visual wavelengths. This result has been confirmed by subsequent studies (\\eg Wilson and Merrill 1942, Osvalds and Risley 1961, Clayton and Feast 1969), however the scatter of this period--luminosity dependence turned out to be very large. First tight {\\it PL} relation for LPVs was discovered for Mira stars at near infrared (NIR) wavebands (Glass and Lloyd Evans 1981). This {\\it PL} law, based on only 11 Miras in the LMC, was refined by extensive studies of Feast \\etal (1989) and Hughes and Wood (1990). The second, parallel {\\it PL} sequence, occupied by semiregular variables (SRVs), was identified by Wood and Sebo (1996). This sequence was shifted relative to the Miras' ridge toward shorter periods by a factor of two. However, the subject of {\\it PL} distribution of LPVs has progressed rapidly over recent years when large microlensing surveys (MACHO, OGLE, EROS, MOA) published long-term photometry of huge number of stars. Complex structure of the {\\it PL} distribution was demonstrated for the first time by Cook \\etal (1997), who published {\\it PL} diagram for variable stars detected during the MACHO survey in the LMC. A series of three or four {\\it PL} sequences defined by LPVs can be distinguished in that diagram. Sharper picture was presented by Wood \\etal (1999), who distinguished and described five {\\it PL} sequences (denoted as A--E) in the period--Wesenheit index plane. Wood (2000) showed similar distribution in the $\\log P$--$K$ diagram. These results were then confirmed by many studies based on observations originated in various sources (Cioni \\etal 2001, 2003, Noda \\etal 2002, Lebzelter \\etal 2002, Ita \\etal 2004, Groenewegen 2004, Fraser \\etal 2005). Kiss and Bedding (2003, 2004) used OGLE data to reveal new features in the {\\it PL} distribution. They noticed that sequence B consists of two closely spaced parallel ridges (Ita \\etal 2004 denoted the additional sequence as C$'$). Below the tip of the red giant branch (TRGB) Kiss and Bedding (2003) found three sequences shifted in $\\log P$ relative to stars brighter than TRGB. It was the definitive proof that stars in the first ascent Red Giant Branch (RGB) pulsate similarly to objects being in the Asymptotic Giant Branch (AGB) phase. The Optical Gravitational Lensing Experiment (OGLE) collected unprecedented amount of photometric data of stars in the Large and Small Magellanic Clouds. Both galaxies have been constantly monitored since 1997 and at present time this is the best and longest available photometric dataset for analyzing huge number of variable red giants. Our studies on LPVs resulted in many discoveries, regarding also the {\\it PL} relations. Soszy{\\'n}ski \\etal (2004a) showed that OGLE Small Amplitude Red Giants (OS\\-ARGs) constitute separate class of variable stars, with different structure in the {\\it PL} plane than ``classical'' SRVs and Miras. We indicated two previously overlooked {\\it PL} relations -- the longest (${\\rm a}_1$) and the shortest (${\\rm a}_4$) period sequences followed by AGB OSARGs. We also suggested a method of empirical division between RGB and AGB OSARGs fainter than TRGB. Red giants revealing ellipsoidal modulation caused by binarity were analyzed by Soszy{\\'n}ski \\etal (2004b). It was shown that, if true orbital periods are considered, the {\\it PL} relation of ellipsoidal variables (sequence~E) is a direct continuation of sequence~D occupied by mysterious Long Secondary Period (LSP) variables. This is a hint that the LSP phenomenon may be related to binarity, but taking into account available radial velocity measurements, the secondary component usually must be a low mass object, possibly former planet. This idea was supported by Soszy{\\'n}ski (2007) who discovered in some LSP variables ellipsoidal-like and eclipsing-like modulations with periods equal to LSPs. In Soszy{\\'n}ski \\etal (2005) we again increased complexity of the PL distribution of LPVs. Each of the NIR {\\it PL} sequences C$'$, C and D in the LMC (occupied by SRV, Miras and LSP variables) split into two separate ridges in the period -- optical Wesenheit index plane, what corresponds to the spectral division into oxygen-rich (O-rich) and carbon-rich (C-rich) AGB stars. Thus, we found a new photometric method of distinguishing between these two populations. In this paper we describe in details the {\\it PL} relations of variable red giants in both Magellanic Clouds. We show new details in the {\\it PL} plane and compare {\\it PL} distribution in the LMC and SMC. The paper is organized as follows. Section~2 gives details of the observations and data reduction. In Section~3 the {\\it PL} relations are presented with a description of their derivation. A discussion about four types of red giant variability -- OSARGs, Miras/SRVs, LSPs and ellipsoidal modulation -- is given in Sections~4--7. Section~8 summarizes and concludes the paper. ", "conclusions": "In this paper we showed the most complex structure of the {\\it PL} distribution presented so far. The Wood's five ridges turn out to be an overlap of fourteen sequences (if consider closely spaced {\\it PL} relations of OSARGs, the number of sequences exceeds twenty). In order to help recognizing {\\it PL} relations with published {\\it PL} laws we provide Table~3 with appropriate identifications. Diagrams employing period ratios (similar to the Petersen diagram) were used as a tool for discriminating OGLE Small Amplitude Red Giants (OSARGs). We compared three sequences of the {\\it PL} relations for the LMC OSARGs in the RGB phase with the calculated relation for first three radial modes using isochrone calculations. We found an essential agreement with our knowledge about metallicities and ages of red giants population in the LMC. However, there are also discrepancies which may suggest the necessity of refinement of the knowledge and/or stellar models. This is a potential application of the OSARGs. \\MakeTableTop{l@{\\hspace{4pt}}c@{\\hspace{4pt}} l@{\\hspace{8pt}} c@{\\hspace{8pt}} c@{\\hspace{8pt}} c@{\\hspace{8pt}}}{12.5cm} {Labels of the {\\it PL} relations in this and previous papers} {\\hline \\noalign{\\vskip3pt} \\multicolumn{3}{c}{this paper} & Wood \\etal 1999 & Kiss and Bedding 2003 & Ita \\etal 2004 \\\\ \\noalign{\\vskip3pt} \\hline \\noalign{\\vskip3pt} & & ${\\rm b}_1$ & & $R_1$ & \\\\ & RGB & ${\\rm b}_2$ & B & $R_2$ & $B^-$ \\\\ & & ${\\rm b}_3$ & A & $R_3$ & $A^-$ \\\\ OSARGs & & ${\\rm a}_1$ & & & \\\\ & & ${\\rm a}_2$ & B & 2O & $B^+$ \\\\[-1ex] & \\raisebox{1.5ex}{AGB} & ${\\rm a}_3$ & A & 3O & $A^+$ \\\\ & & ${\\rm a}_4$ & & & \\\\ \\noalign{\\vskip3pt} \\hline \\noalign{\\vskip3pt} & & C$_{\\rm O}$ & C & F & C \\\\[-1ex] Miras & \\raisebox{1.5ex}{O-rich} & C$'_{\\rm O}$ & B & 1O & C$'$ \\\\ and SRVs & & C$_{\\rm C}$ & C & F & C \\\\[-1ex] & \\raisebox{1.5ex}{C-rich} & C$'_{\\rm C}$ & B & 1O & C$'$ \\\\ \\noalign{\\vskip3pt} \\hline \\noalign{\\vskip3pt} & O-rich & D$_{\\rm O}$ & D & $L_2$ & D \\\\[-1ex] \\raisebox{1.5ex}{LSPs} & C-rich & D$_{\\rm C}$ & D & $L_2$ & D \\\\ \\noalign{\\vskip3pt} \\hline \\noalign{\\vskip3pt} \\multicolumn{2}{l}{Ell. and Ecl.} & E & E$^*$ & $L_1^*$ & E$^*$ \\\\ \\noalign{\\vskip3pt} \\hline \\multicolumn{6}{l}{$^*$ -- the sequence is shifted due to halving the orbital periods.} } Most likely mechanism responsible for pulsation in the OSARGs is the stochastic excitation. Therefore we looked at these objects as solar-like pulsators. The range of periods agrees with predictions based on extrapolation of the relations found for much less luminous stars. The amplitudes are lower than predicted by the Kjeldsen and Bedding (1995) formula, where the the amplitude rises linearly with the luminosity-to-mass ratio. However, there are calculations predicting a slower amplitude rise. Testing theory of stochastic excitation is another possible application of OSARGs. We would like to bring the reader's attention to NIR Wesenheit index ($W_{JK}$). The sequences in the period--$W_{JK}$ plane are generally better defined than those at $K_S$ magnitudes. Wesenheit index is a reddening independent quantity, so even heavy reddened Miras fall on the {\\it PL} sequence. Therefore, the period--$W_{JK}$ relations can be used as a distance indicators without correcting them for reddening. Moreover, O- and C-rich Miras and SRV obey very similar relations in the period--$W_{JK}$ plane, while at $K_S$ the {\\it PL} relations are significantly different for both populations. We conclude that $W_{JK}$ index can be a useful tool for studying LPVs. Comparison of the {\\it PL} relations in NIR supports the binary star scenario as the explanation of the LSP variability. We further develop this scenario taking into account data on the ratios of the short period (pulsational) and LSP variability. For OSARGs these ratio is nearly constant which translates into nearly constant ratio of the stellar to orbital radius the value of about 0.4. We proposed that the small mass companion position is determined by the balance between mass loss and tidal effects. This scenario requires that the proximity of the companion enhances mass loss. Moreover, it requires that a substantial fraction (majority) of red giants have such companions and that their mass cannot be too small. \\vspace*{9pt} \\Acknow{The paper was supported by the Foundation for Polish Science through the Homing Program and by MNiSW grants: 1P03D01130 and N20303032/4275. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation.} \\vspace*{9pt}" }, "0710/0710.5505_arXiv.txt": { "abstract": "We analyze a toy swiss-cheese cosmological model to study the averaging problem. In our swiss-cheese model, the cheese is a spatially flat, matter only, Friedmann-Robertson-Walker solution (\\textit{i.e.,} the Einstein--de Sitter model), and the holes are constructed from a Lema\\^{\\i}tre-Tolman-Bondi solution of Einstein's equations. We study the propagation of photons in the swiss-cheese model, and find a phenomenological homogeneous model to describe observables. Following a fitting procedure based on light-cone averages, we find that the the expansion scalar is unaffected by the inhomogeneities (\\textit{i.e.,} the phenomenological homogeneous model is the cheese model). This is because of the spherical symmetry of the model; it is unclear whether the expansion scalar will be affected by non-spherical voids. However, the light-cone average of the density as a function of redshift is affected by inhomogeneities. The effect arises because, as the universe evolves, a photon spends more and more time in the (large) voids than in the (thin) high-density structures. The phenomenological homogeneous model describing the light-cone average of the density is similar to the $\\Lambda$CDM concordance model. It is interesting that although the sole source in the swiss-cheese model is matter, the phenomenological homogeneous model behaves as if it has a dark-energy component. Finally, we study how the equation of state of the phenomenological homogeneous model depends on the size of the inhomogeneities, and find that the equation-of-state parameters $w_{0}$ and $w_{a}$ follow a power-law dependence with a scaling exponent equal to unity. That is, the equation of state depends linearly on the distance the photon travels through voids. We conclude that within our toy model, the holes must have a present size of about $250$ Mpc to be able to mimic the concordance model. ", "introduction": "Most, if not all, observations are consistent with the cosmic concordance model, according to which one-fourth of the present mass-energy of the universe is clustered and dominated by cold dark matter (CDM). The remaining three-quarters is uniform and dominated by a fluid with negative pressure (dark energy, or $\\Lambda$). While the standard $\\Lambda$CDM model seems capable of accounting for the observations, 95\\% of the mass-energy of the present universe is unknown. This is either a feature, and we are presented with the opportunity of discovering the nature of dark matter and dark energy, or it is a bug, and nature might be different than described by the $\\Lambda$CDM model. Regardless, until such time as dark matter and dark energy are completely understood, it is useful to look for alternative cosmological models that fit the data. One non-standard possibility is that there are large effects on the {\\it observed} expansion rate (and hence on other observables) due to the back-reaction of inhomogeneities in the universe (see, \\textit{e.g.,} Ref.\\ \\cite{kmr,notari,rasa, buchert} and references therein). The basic idea is that all evidence for dark energy comes from the observational determinations of the expansion history of the universe. Anything that affects the observed expansion history of the universe alters the determination of the parameters of dark energy; in the extreme it may remove the need for dark energy. The ``safe'' consequence of the success of the concordance model is that the isotropic and homogeneous $\\Lambda$CDM model is a good {\\it phenomenological} fit to the real inhomogeneous universe. And this is, in some sense, a verification of the cosmological principle: the inhomogeneous universe can be described by means of an isotropic and homogeneous solution. However, this does not imply that a primary source of dark energy exists, but only that it exists as far as the phenomenological fit is concerned. For example, it is not straightforward that the universe is accelerating. If dark energy does not exist at a fundamental level, its presence in the concordance model would tell us that the pure-matter inhomogeneous model has been renormalized, from the phenomenological point of view (luminosity-distance and redshift of photons), into a homogeneous $\\Lambda$CDM model. The issue is the observational significance of the back-reaction of inhomogeneities. Our point of view is tied to our past light cone: we focus on the effects of large-scale nonlinear inhomogeneities on observables such as the luminosity-distance--redshift relation. We will not discuss averaged domain dynamics, even though if we think it is a crucial step in understanding how General Relativity effectively works in a lumpy universe \\cite{ellis, buchert_new}. Following this approach, we built in Ref.\\ \\cite{marra-sc} a particular swiss-cheese model, where the cheese consists of a spatially flat, matter only Friedmann-Robertson-Walker (FRW) solution and the holes are constructed out of a Lema\\^{i}tre-Tolman-Bondi (LTB) solution of Einstein's equations. We attempted to find a model that was solvable and ``realistic'' (even if still toy), rather than finding a model with interesting volume-averaged dynamics. The model, however, will turn out to be useful to investigate light-cone averages. It has been indeed shown that the LTB solution can be used to fit the observed $d_{L}(z)$ without the need of dark energy (for example, see Ref.\\ \\cite{alnes0602}). To achieve this result, however, it is necessary to place the observer at the center of a rather large-scale underdensity. To overcome this fine-tuning problem we built a swiss-cheese model with the observer in the cheese looking through a series of holes. In Ref.\\ \\cite{marra-sc} we studied this model in detail and discussed the effects of large-scale nonlinear inhomogeneities on observables such as the luminosity-distance--redshift relation. We found that inhomogeneities are able (at least partly) to mimic the effects of dark energy. In this paper we will analyze the same swiss-cheese model through the fitting scheme developed by Ellis and Stoeger \\cite{ellis-f} in order to better understand how inhomogeneities renormalize the (matter only) swiss-cheese model allowing us to avoid a physical dark-energy component. We think that this model fits well in that context and therefore we might be able to shed some light on the important topics discussed there. We will propose a fitting procedure based on light-cone averages. The paper is organized as follows: In Sec.\\ \\ref{model} we will specify the parameters of our swiss-cheese model and summarize the main results obtained in Ref.\\ \\cite{marra-sc}. In Sec.\\ \\ref{fitti}, we develop our fitting procedure, and in Sec.\\ \\ref{disco} we discuss our results. Then, in Sec.\\ \\ref{dressing} we study the dependence of the best-fit parameters on the size of the holes. Conclusions are given in Sec.\\ \\ref{conclusions}. ", "conclusions": "\\label{conclusions} The aim of this investigation was to understand the role of large-scale non-linear cosmic inhomogeneities in the interpretation of observational data. We focused on an exact (if toy) solution, based on the Lema\\^{\\i}tre-Tolman-Bondi (LTB) model. This solution has been studied extensively in the literature \\cite{alnes0607, notari-mansouri, alnes0602, celerier, mansouri, flanagan, rasanen, tomita, chung, nambu}. It has been shown that it can be used to fit the observed $d_{L}(z)$ without the need of dark energy (for example in Ref.\\ \\cite{alnes0602}). To achieve this result, however, it is necessary to place the observer at the center of a rather large-scale underdensity. To overcome this fine-tuning problem we built a swiss-cheese model, placing the observer in the cheese and having the observer look through the holes in the swiss-cheese as pictured in Fig.\\ \\ref{schizzo}. In Sec.\\ \\ref{model} we defined the model and described its dynamics: it is a swiss-cheese model where the cheese is made of the usual FRW solution and the holes are made of a LTB solution. The voids inside the holes are expanding faster than the cheese. We reported also the results for $d_{L}(z)$ obtained in Ref.\\ \\cite{marra-sc}, to which we refer the reader for a more thorough analysis. We found that redshift effects are suppressed because of a compensation effect due to spherical symmetry. However, we found interesting effects in the calculation of the angular distance: the evolution of the inhomogeneities bends the photon path compared to the FRW case. Therefore, inhomogeneities will be able (at least partly) to mimic the effects of dark energy. After having analyzed the model from the observational point of view, we set up in Section \\ref{fitti} the fitting problem in order to better understand how inhomogeneities renormalize the matter swiss-cheese model allowing us to eschew a primary dark energy. We followed the scheme developed in Ref.\\ \\cite{ellis-f}, but modified in the way to fit the phenomenological model to the swiss-cheese one. We chose a method that is intermediate between the fitting approach and the averaging one: we fitted with respect to light-cone averages. In particular, we focused on the expansion and the density. While the expansion behaved as in the FRW case because of the compensation effect mentioned above, we found that the density behaved differently thanks to its intensiveness to that compensation effect: a photon is spending more and more time in the (large) voids than in the (thin) high density structures. This effect is not directly linked to the one giving us an interesting $d_{A}$. The best fit we found for holes of $r_{h}=250$ Mpc is $w_{0}=-1.03$ and $w_{a}=2.19$; qualitatively similar to the concordance model. The flow chart of Fig.\\ \\ref{schema} summarizes the results obtained. The insensitivity to the compensation effect made us think that a swiss cheese made of spherical symmetric holes and a swiss cheese without an exact spherical symmetry would share the same light-cone averaged density. Knowing the behavior of the density we are therefore able to know the one of the Hubble parameter that will be the one of the FRW solution with an phenomenological source characterized by the fit equation of state. In this way we can think to go beyond the main limitation of this model, that is, the assumption of spherical symmetry. From this point of view, the light-cone averaged density can be seen as a tool in performing this step. Summarizing: \\begin{itemize} \\item We started with a swiss-cheese model based on spherically symmetric holes. A photon, during its journey through the swiss cheese, undergoes a redshift which is not affected by inhomogeneities. However the photon is spending more and more time in the voids than in the structures. The lack of an effect is due to the the assumption of spherical symmetry. We focus on this because a photon spending most of its time in voids should have a different redshift history than a photon propagating in a homogeneous background. \\item Assuming that the density is a quantity that does not heavily depend on the assumption of spherical symmetry, we tried to resolve the issue by focusing on the density alone and getting from it the expansion (and therefore the redshift history). \\item This resulted in a swiss-cheese model with holes that effectively are not perfectly spherical. In this model the redshift history of a photon depends on the time passed inside the voids. \\item In practice this means that we will use the phenomenological best-fit model found, that is, we will use a model that behaves similarly to the concordance model. \\end{itemize} Then, in Section \\ref{dressing} we studied how the equation of state of a phenomenological model with only one effective source depends on the size of the inhomogeneity. We found that $w^{R}_{0}$ and $w^{R}_{a}$ follow a power-law dependence with the same scaling exponent which is equal to unity. That is, the equation of state depends linearly on the distance the photon travels through voids. We finally asked which size of the holes will give us a phenomenological model able to mimic the concordance model. We found that for $n=1$, that is for a holes of radius $r_{h}=250$ Mpc, we have $w_{0}=-1.03$ and $w_{a}=2.19$." }, "0710/0710.5219_arXiv.txt": { "abstract": "{} {The analysis of near infrared spectropolarimetric data at the internetwork at different regions on the solar surface could offer constraints to reject current modeling of these quiet areas.} {We present spectro-polarimetric observations of very quiet regions for different values of the heliocentric angle for the Fe\\,{\\sc i} lines at $1.56$ $\\mu$m, from disc centre to positions close to the limb. The spatial resolution of the data is $0.7-1''$. We analyze direct observable properties of the Stokes profiles as the amplitude of circular and linear polarization as well as the total degree of polarization. Also the area and amplitude asymmetries are studied.} {We do not find any significant variation of the properties of the polarimetric signals with the heliocentric angle. This means that the magnetism of the solar internetwork remains the same regardless of the position on the solar disc. This observational fact discards the possibility of modeling the internetwork as a Network-like scenario. The magnetic elements of internetwork areas seem to be isotropically distributed when observed at our spatial resolution.} {} ", "introduction": "The presence of magnetic fields in the solar internetwork was discovered more than 30 years ago \\citep{livingston_75,smithson_75}. Since that work, improvements on the instrumentation have allowed the observation of the full Stokes vector in those regions with a signal-to-noise ratio good enough to retrieve the magnetic field vector from the observational data. Nevertheless, the internetwork magnetic topology remains yet very indistinct and shows a growing complexity as we improve the quality of the data. \\cite{martin_87} observed that in her videomagnetographs at a spatial resolution of $3''$ the longitudinal component of internetwork magnetic fields was present everywhere in the solar disc. She immediately concluded that these magnetic structures should have very small scales and a very tangled geometry, giving as an example a scenario in which the magnetic fields in the internetwork would consist of a maze of small loops. As spectro-polarimeters have made possible the precise detection of these signals, the efforts have concentrated on the study of the distribution of magnetic fields at disc centre. Different methodologies (mainly based upon the use of the Zeeman and Hanle effects) shed light on different aspects of the internetwork magnetism. Concerning the Zeeman effect, the Stokes $Q$, $U$ and $V$ signals detected on the most widely used spectral lines (Fe\\,{\\sc i} at $1.5$ $\\mu$m and 630 nm) are very weak at the best spatial resolutions of 0.5-1$''$. Moreover, the Stokes $I$ profile seems to come from a field free atmosphere, as expected from weakly polarized media \\citep{jorge_99}. The filling factor of the magnetic elements retrieved from the analysis of these signals is always around 2 \\% \\citep{khomenko_03, jorge_ita_03, marian_spw4}. The rest of the resolution element would be filled with very weak magnetic fields. Another possibility is that the real element is filled with mixed polarity magnetic fields which would partially cancel out due to the lack of spatial resolution. The magnetic field strength distributions recovered for this 2 \\% of the resolution element appear to show a preference for magnetic fields around the equipartition field at photospheric heights and even weaker \\citep{khomenko_03, marian_spw4, julio07}. Additionally, \\cite{andres_07} have shown the first direct observational evidence of flux cancellation in the internetwork (note that their internetwork region is surrounded by a very enhanced network structure), showing that more than 95 \\% of the magnetic flux is cancelled in the resolution element ($\\sim 1''$). This means that the above-mentioned distributions could not give a complete vision of the internetwork magnetism. \\cite{andres_07} give an amount of $250$ G for the mean magnetic field in the resolution element (note that the values of the magnetic flux density obtained by means of the Zeeman effect are below 10 Mx/cm$^2$). This would mean that the magnetism of the internetwork could play an important role on the solar global magnetism. This has also been pointed out by works using the Hanle effect \\citep{javier_04}. The study of regions in different positions of the solar disc could represent a strong constraint to reject models for the solar internetwork. Using the 630 nm lines with a spatial resolution of $\\sim 1''$, \\cite{lites_02} built a histogram of both circular and linear polarization signals in two quiet regions, one at disc centre and another at $\\mu=0.82$ (being $\\mu$ the cosine of the heliocentric angle). He did not observe significant linear polarization signals in neither of the two regions. However, Figure 9 in \\cite{lites_02} shows that the histograms of circular polarization do not present any variation for those signals whose integrated signal is below 0.005. \\cite{meunier_98}, studying integrated polarization, and \\cite{harvey_07} using magnetograms, show the presence of a horizontal component of the magnetic field everywhere in the solar disc. In this paper we present the first study of the internetwork at several positions on the solar disc using high quality $0.8''$ spectro-polarimetric data. ", "conclusions": "We have analyzed high quality spectro-polarimetric data of the Fe\\,{\\sc i} lines at $1.56$ $\\mu$m in order to infer information about the internetwork magnetism at different positions on the Sun's surface. The whole field of view presents significant signal, meaning that the magnetic fields pervade the observed areas. We have found that the circular and linear polarization amplitudes do not have any clear dependence on the heliocentric angle. This fact goes against a Network-like scenario for the internetwork: quasi--vertical flux tubes cannot explain this observational result, nor in fact any field topology with a preferred orientation within the field-of-view. An isotropical distribution of magnetic fields, oriented in all directions in the whole field of view, is on the other hand expected to show this behaviour. \\cite{marian_07} found that at least 10-20 \\% of the magnetic flux in the internetwork is connected by low-lying loops. Consequently, the scenario proposed by \\cite{martin_87} of an internetwork characterized by a myriad of small loops is a very reasonably idea that is compatible with all the observational constraints presented in this work. Of course one can think of other scenarios that are compatible with the observations. \\citep{stenflo_87, rafa_04, javier_04} adopt turbulent internetwork magnetism, \\cite{jorge_00} proposes a micro-structuration of the atmosphere (MISMA) to explain all the magnetic phenomena on the solar surface. All of them are compatible with the presented results and our efforts should be headed towards finding more constraints to reject some of them and strengthen others. The size of the magnetic structures can also be constrained by the information presented in this study. First, improving the spatial resolution, we do not see a global increase in the signals. \\cite{lites_04} found no increment of the magnetic flux density from $1\"$ to $0.6\"$. Recently \\cite{lites_07} computed a magnetic flux density of about $11$ Mx/cm$^2$ at 0.3$''$ using HINODE's data, which is compatible with the value of $10$ Mx/cm$^2$ found by \\cite{martin_87} at $3''$. This means that, either the magnetic field structures are already resolved at $\\sim 0.5''$ or we are very far from resolving them. The fact that the polarimetric signals do not vary along the solar surface would point towards very small structures as the responsibles for the internetwork magnetism. The size of these magnetic structures is something not yet constrained by the observations. We have presented a study of high quality spectro-polarimetric data in different positions of the solar surface, from the disc centre towards $\\mu=0.28$. This is the first step of the study of the variation on the magnetism of the internetwork with the heliocentric angle. Much more work has to be done by retrieving physical parameters as the magnetic field strength vector, magnetic flux, etc. to really constrain the modeling of the internetwork and reject models that are not compatible with the results." }, "0710/0710.4134_arXiv.txt": { "abstract": "Since 1999, a radial velocity survey of 179 red giant stars is ongoing at Lick Observatory with a one month cadence. At present $\\sim$20$-$100 measurements have been collected per star with an accuracy of 5 to 8 m\\,s$^{-1}$. Of the stars monitored, 145 (80\\%) show radial velocity (RV) variations at a level $>$20 m\\,s$^{-1}$, of which 43 exhibit significant periodicities. Here, we investigate the mechanism causing the observed radial velocity variations. Firstly, we search for a correlation between the radial velocity amplitude and an intrinsic parameter of the star, in this case surface gravity ($\\log g$). Secondly, we investigate line profile variations and compare these with theoretical predictions. ", "introduction": "Since 1999, a radial velocity survey of 179 red giant stars is ongoing at Lick Observatory, using the 60 cm Coud\\'e Auxiliary Telescope (CAT) in conjunction with the Hamilton echelle spectrograph (R $\\approx$ 60\\,000). These stars have been selected from the Hipparcos catalogue \\cite{esa1997}, based on the criteria described by \\cite{frink2001}. The selected stars are all brighter than 6~mag, are presumably single and have photometric variations $< 0.06$~mag in V. The system with an iodine cell in the light path has been developed as described by \\cite{marcy1992} and \\cite{valenti1995}. With integration times of up to thirty minutes for the faintest stars ($m_{v}$ = 6 mag) we reach a signal to noise ratio of about $80-100$ at $\\lambda = 5500$ \\AA , yielding a radial velocity precision of $5-8$ m\\,s$^{-1}$. The initial aim of the survey was to check whether red giants would be stable enough to serve as reference stars for astrometric observations with SIM/PlanetQuest \\cite{frink2001}. In \\cite{hekker2006} it is shown that a large fraction of the red giants in a specific part of the absolute magnitude vs. B-V colour diagram are stable to a level of 20 m\\,s$^{-1}$ and could be effectively searched for long period companions, as is required for astrometric reference stars. For other stars in the sample the radial velocity variations are larger, and for 43 stars these show significant periodicities. So far, sub-stellar companions have been announced for two stars from the present sample ($\\iota$ Dra \\cite{frink2002} and $\\beta$ Gem \\cite{reffert2006}). Here, we investigate which physical mechanism causes the observed radial velocity variations. In cases for which we do not find a significant periodicity in the observed radial velocity variations, an intrinsic mechanism such as spots or pulsations, possibly multi-periodic, seems most likely. On the other hand, the periodic radial velocity variations can be caused by sub-stellar companion, an intrinsic mechanism, or by both these mechanisms simultaneously. In Section 2 we search for a relation between the amplitude of the radial velocity variations and an intrinsic parameter, i.e.~$\\log g$. In Section 3 we investigate line shape variations and compare these with theoretical predictions. Our conclusions are presented in Section 4. A more extended paper on this subject is submitted \\cite{hekker2007b}. ", "conclusions": "There exists a clear correlation between the half peak-to-peak values of the radial velocity variations and the surface gravity of the red giant stars. This is a strong indication that the observed radial velocity variations are caused by a mechanism intrinsic to the star. Companions and an intrinsic mechanism might be present in stars with periodic radial velocity variations and $\\log g > 1.6$ dex, while for stars with $\\log g<1.6$ dex solely an intrinsic mechanism seems most likely. We have investigated whether there is a relation between the radial velocity amplitude and the amplitude of the line depth residuals. From theory we find that the rotational velocity, intrinsic and equivalent line width largely influence the line depth residual and no clear correlation could be identified. In the theoretical models temperature variations due to the pulsations is ignored, but these temperature variations influence the intrinsic and equivalent line width of the spectral line, which influences the line depth residuals. This might be an explanation why the theoretical models do not overlap with the observations." }, "0710/0710.4072_arXiv.txt": { "abstract": "{We investigate correlations between the optical linear polarization position angle and the orientation of the host galaxy/extended emission of Type~1 and Type~2 Radio-Loud (RL) and Radio-Quiet (RQ) quasars. We have used high resolution Hubble Space Telescope ({\\it HST}) data and deconvolution process to obtain a good determination of the host galaxy orientation. With these new measurements and a compilation of data from the literature, we find a significant correlation between the polarization position angle and the position angle of the major axis of the host galaxy/extended emission. The correlation appears different for Type~1 and Type~2 objects and depends on the redshift of the source. Interpretations in the framework of the unification model are discussed.} \\addkeyword{Quasars : general} \\addkeyword{Polarization} \\begin{document} ", "introduction": " ", "conclusions": "We can summarize our results as follow : \\begin{asparaitem} \\item{} While Type~2 RL and RQ quasars are known to exhibit an anti-alignment between the major axis of hteir host/extended emission and their optical polarization, we find that an alignment is mostly observed for Type~1 quasars. \\item{} The redshift dependence of the alignment effect, and lack of correlation with the near-IR $PA_{host}$, suggest that it might be related to the rest-frame extended UV/blue emission of quasars. \\item{}We show that these observations can be interpreted in the framework of the unification model + a two component scattering model. \\end{asparaitem}" }, "0710/0710.3378_arXiv.txt": { "abstract": "In order to prepare the analysis of COROT data, it has been decided to build a simple tool to simulate the expected light curves. This simulation tools takes into account both instrumental constraints and astrophysical inputs for the COROT targets. For example, granulation and magnetic activity signatures are simulated, as well as p modes, with the expected photon noise. However, the simulations rely sometimes on simple approach of these phenomenons, as the main goal of this tool is to prepare the analysis in the case of COROT data and not to perform the most realistic simulations of the different phenomenons. ", "introduction": "Simulating the data that a space instrument like COROT will provide might look presomptuous. Indeed, it is certainly, when comparing to previous comparable instruments like IPHIR or GOLF. These two examples show that the nominal behaviour of the instrument is not always reached, but this does not prevent this instrument to provide very interesting data. However, despite some technical problems, IPHIR and GOLF yielded a wealth of scientific results. Thus, what is the interest of simulating COROT data? How close to reality these simualtions will get? This might not be the most important fact as the preparation of these simulations will help us to prepare the analysis of real data and to be ready in case of unexpected technical behaviour of the instrument perturbating the data, or unexpected physical behaviour of the targets of the instrument. A consequence of that is that the simulation tool must include technical and physical aspects, making the task even more difficult. These aspects cover: photon noise, p modes excitation, granulation signal, stellar activity signal, orbital perturbations, stellar rotation... The software presented here is freely available at:\\\\ {\\tt www.lesia.obspm.fr/$\\sim$corotswg/simulightcurve.html} ", "conclusions": "This simulator software will continue to evolve with time. As indicated above, intensity modulation due to starspots will be included, as well as other stellar or instrumental signals, as for example instrumental perturbations due to orbital vraiations. Moreover, this effort of simulation will not end with the delivery of first data but will be continued after that. The comparison with real data will allow to check for the validity of physical hypothesis used to simulate the different signals of astrophysics origin in the data. This shoud bring a great amount of information on our knowmedge of these often not well known phenomena, which stellar simulation is often derived from the solar case. In parallel, the simulation of instrumental components of the signal will be improved to help the interpretation of real data. All these reasons justify in our opinion the need for the simulation tool presented here." }, "0710/0710.4244_arXiv.txt": { "abstract": "We summarise the mathematical foundation of the holographic method of measuring the reflector profile of an antenna or radio telescope. In particular, we treat the case, where the signal source is located at a finite distance from the antenna under test, necessitating the inclusion of the so-called Fresnel field terms in the radiation integrals. We assume a ``full phase'' system with reference receiver to provide the reference phase. We describe in some detail the hardware and software implementation of the system used for the holographic measurement of the 12m ALMA prototype submillimeter antennas. We include a description of the practicalities of a measurement and surface setting. The results for both the VertexRSI and AEC (Alcatel-EIE-Consortium) prototype ALMA antennas are presented. ", "introduction": "\\label{intro} \\PARstart{L}{arge} reflector antennas, as those used in radio astronomy and deep-space communication, generally are composed of a set of surface panels, supported on three or more points by a support structure, often called the backup structure. After assembly of the reflector it is necessary to accurately locate the panels onto the prescribed paraboloidal surface in order to obtain the maximum antenna gain. The fact that some antennas have a \"shaped\"contour is irrelevant for the purpose of our discussion. We are concerned with describing a method which allows us to derive the position of the individual panels in space and compute the necessary adjustments of their support points to obtain a continuous surface of a certain prescribed shape. The analysis by Ruze \\cite{Ruze1966} of the influence of random errors in the reflector contour on the antenna gain indicates that the RMS error should be less than about one-sixteenth of the wavelength for acceptable performance. Under the assumption that the errors are small compared to a wavelength, randomly distributed with RMS value \\(\\epsilon \\), have a correlation length {\\bfseries c} which is much larger than the wavelength \\(\\lambda \\), and much smaller than the reflector diameter D, the relative decrease in aperture efficiency (or gain) can be expressed by the simple formula \\begin{equation} \\frac{\\eta_A}{\\eta_{A0}} = \\exp\\left\\{-{\\left(\\frac{4\\pi\\epsilon}{\\lambda}\\right)}^2\\right\\}, \\label{eq:etaa} \\end{equation} \\noindent{where} \\({{\\eta }_{A0}}\\) is the aperture efficiency of the perfect reflector. An error $\\epsilon$ of \\(\\lambda \\)/40 is required to limit the gain loss to 10 percent; with an error of \\(\\lambda \\)/16 the gain is decreased to about half of the maximum achievable. The setting of the reflector panels at accuracies better than 100 \\(\\mu \\)m has required the development of measuring methods of hitherto unsurpassed accuracy. It should be noted that these measurements need to be done ``in the field'', which in the case of millimeter radio telescopes generally means the hostile environment of a high mountain site. One versatile, and by now widely used method is normally called ``radio holography''. The method makes use of a well-known relationship in antenna theory: the far-field radiation pattern of a reflector antenna is the Fourier Transformation of the field distribution in the aperture plane of the antenna. Note that this relationship applies to the amplitude/phase distributions, not to the power pattern. Thus, if we can measure the radiation pattern, in amplitude and phase, we can derive by Fourier Transformation the amplitude and phase distribution in the antenna aperture plane with an acceptable spatial resolution. Bennett \\etal \\cite{Bennett1976} presented a sufficiently detailed analysis of this method to draw the attention of radio astronomers. Scott \\& Ryle \\cite{Scott1977} used the new Cambridge 5 km array to measure the shape of four of the eight antennas, using a celestial radio point source and the remaining antennas to provide the reference signal. Simulation algorithms were developed by Rahmat-Samii \\cite{RahmatSamii1985} and others, adding to the practicability of the method. Using the giant water vapour maser at 22 GHz in Orion as a source Morris \\etal \\cite{Morris1988} achieved a measurement accuracy of 30 $\\mu$m and were able to set the surface of the IRAM 30-m millimeter telescope to an accuracy of better than 100 $\\mu$m. \\pubidadjcol Artificial satellites, radiating a beacon signal at a fixed frequency have also been used as farfield (${R_f} = \\frac{2 {D^2}}{\\lambda}$) signal sources. Extensive use has been made of synchronous communication satellites in the 11 GHz band \\cite{Godwin1986}, \\cite{Rochblatt1992}. These transmitters of course do not provide the range of elevation angles accessible with cosmic sources. Some satellites, notably the LES (Lincoln Experimental Satellite) 8 and 9, have been used for radio holography of millimeter telescopes \\cite{Baars1999}. They provided a signal at the high frequency of 37 GHz and with their geo-synchronous orbit moved over some 60 degrees in elevation angle. Unfortunately, both satellites are no longer available. Lacking a sufficiently strong source in the farfield, we have to take recourse to using an earth-bound transmitter. In practice these will be located at a distance of several hundreds of meters to a few kilometers and be at an elevation angle of less than 10 degrees. Clearly, these are in the nearfield of the antenna, requiring significant corrections to the received signals. In particular, the phase front of the incoming waves will not be plane and it contains higher order terms in the radial coordinate of the antenna aperture. These must be corrected before the Fourier transformation can be applied. We treat these corrections in detail in this paper. Successful measurements on short ranges have been reported for the University of Texas millimeter telescope \\cite{Mayer1983}, the IRAM 30-m telecope \\cite{Morris1988b}, the JCMT \\cite{Hills2002} and the ASTE antenna of NAOJ \\cite{Ezawa2000}. ALMA (Atacama Large Millimeter Array) is a new large aperture synthesis array for submillimeter astronomy consisting of 50 high accuracy antennas of 12 m diameter. The instrument is under construction at 5000 m altitude in the Atacama desert of northern Chile. ALMA is a collaboration of North America and Europe with participation of Japan. Two prototype antennas were procured and erected at the site of the Very Large Array of NRAO in New Mexico. The results of an extensive evaluation program of these antennas has been presented by Mangum \\etal \\cite{Mangum2006}. The reflector surface accuracy was specified at 20-25 $\\mu$m, requiring a measurement method with an accuracy of 10 $\\mu$m or better. This was achieved with a near-field holographic system using a transmitter at a wavelength of 3 mm and at a distance of only 315 m from the antennas at an elevation angle of 9 degrees. Here we describe these measurements in some detail. ", "conclusions": "\\label{conclusions} We have successfully performed a holographic measurement and consecutive panel setting of the reflectors of the two ALMA prototype antennas to an accuracy of better than 20 $\\mu$m. Our estimated measurement accuracy is approximately 5 $\\mu $m. The data collection and analysis software packages are easy to use and provide quick results of the measurements, directly usable for a panel adjustment setting. We consider this system suitable for the routine setting of the ALMA production antennas to the goal of 20 $\\mu$m accuracy in an acceptable time span. Modern survey equipment enables contractors to deliver reflectors with an accuracy of 50-60 $\\mu$m without undue cost. Although the holography system can easily start with a much larger error, in the former case it is feasible to reach the specification with only one panel setting based on holography. We note that these measurements, being performed at one elevation angle only, do not provide information on the gravitationally induced deformation as function of elevation angle. In summary: \\begin{enumerate} \\item The holography system has functioned according to specification and has enabled us to measure the surface of the antenna reflector with a repeatability of better than $10\\mu$m. \\item As shown in Figs.~\\ref{fig:holosVertex} and \\ref{fig:holosAEC}, we have set both antenna surfaces to and accuracy of 16-17 $\\mu$m RMS. This will provide an aperture efficiency of about 65 percent of that of a perfect reflector at the highest observing frequency of 950 GHz. \\item The small differences in the surface maps obtained over several days of measurement are consistent with the measurement repeatability and at best marginally significant. If taken at face value, they indicate that the deformations of the reflector under varying wind and temperature influence are fully consistent with, and probably well within, the specification. This excellent behaviour over time is more important than the actual achieved surface setting. We stopped iteration of the settings after having achieved the goal of less than 20 $\\mu$m. \\item Further information on the performance of the ALMA Prototype Antennas can be found in \\cite{Mangum2006}. \\end{enumerate} \\appendices" }, "0710/0710.2654_arXiv.txt": { "abstract": "In this letter we briefly describe the first results of our numerical study on the possibility of magnetic origin of relativistic jets of long duration gamma ray bursters within the collapsar scenario. We track the collapse of massive rotating stars onto a rotating central black hole using axisymmetric general relativistic magnetohydrodynamic code that utilizes a realistic equation of state of stellar matter, takes into account the cooling associated with emission of neutrinos, and the energy losses due to dissociation of nuclei. The neutrino heating is not included. We describe the solution for one particular model where the progenitor star has magnetic field $B=3\\times10^{10}$G. The solution exhibits strong explosion driven by the Poynting-dominated jets whose power exceeds $2\\times10^{51}\\,\\mbox{erg/s}$. The jets originate mainly from the black hole and they are powered via the Blandford-Znajek mechanism. ", "introduction": "\\label{introduction} The phenomenon of Gamma Ray Burst (GRB) has been puzzling astrophysicists for many years since its discovery in 1970s~\\cite{KSO73,MGI74}. The recent identification of long duration GRBs with supernovae (see Della Valle 2006, and Woosley \\& Bloom 2006 for full review) means that we are dealing with enormous amount of energy, $10^{51}-10^{52}\\mbox{erg}$, released within a very short time, 2-100 seconds, in the form of highly relativistic collimated outflow \\cite{P05}. Most of the current GRB studies are focused on the physics associated with production of gamma rays in such flows and their interaction with the interstellar medium or the stellar wind of the supernova progenitor. However, the central question in the problem of GRBs is undoubtedly the nature of their central engines. These powerful jets have to be produced as a result of stellar collapse, most likely by the relativistic object, neutron star or black hole (BH), formed in the center, and make their way through the massive star unscathed, remaining well collimated and highly relativistic. The most popular model of central engine is based on the ``failed supernova'' scenario of stellar collapse, or ``collapsar'', where the iron core of progenitor star forms a BH \\cite{W93}. If the progenitor is non-rotating then its collapse is likely to continue in a ``silent'' manner until the whole star is swallowed by the BH. If, however, the specific angular momentum in the equatorial part of stellar envelope exceeds that of the last stable orbit of the BH then the collapse becomes highly anisotropic. While in the polar region it may proceed more or less uninhibited, for a while, the equatorial layers form dense and massive accretion disk. The gravitational energy released in the disk can be very large, more then sufficient to stop the collapse of outer layers and drive GRB outflows, presumably in the polar direction where density is much lower \\cite{MW99}. In addition, there is plenty of rotational energy in the BH itself \\begin{equation} E\\sub{rot} = \\frac{M\\sub{bh}c^2}{2} \\left\\{ 2-\\left[ \\left(1+\\sqrt{1-a^2}\\right)^2+a^2 \\right]^{1/2} \\right\\}, \\end{equation} where $M\\sub{bh}$ is the BH mass and $a\\in(-1,1)$ is its dimensionless rotation parameter. For $M\\sub{bh}=3M_{\\sun}$ and $a=0.9$ this gives the enormous value of $E\\sub{rot} \\simeq8\\times10^{53}$erg. The three currently actively discussed mechanisms of powering GRB jets in the collapsar scenario are the heating via annihilation of neutrinos produced in the disk \\cite{MW99}, the magnetic braking of the disk \\cite{BP82,UM06}, and the magnetic braking of the BH \\cite{BZ77}. The potential role of neutrino mechanism is rather difficult to assess as this requires accurate treatment of neutrino transport in a complex dynamic environment of collapsar. The long and complicated history of numerical studies of neutrino-driven supernova explosions teaches us to be cautious. Numerical simulations by MacFadyen \\& Woosley\\shortcite{MW99} and Aloy et al.\\shortcite{AIMGM00} have demonstrated that sufficiently large energy deposition in the polar region above the disk may indeed result in fast collimated jets. However, the neutrino transport has not been implemented in these simulations and the energy deposition was based simply on expectations. When Nagataki et al.\\shortcite{NTMT07} utilized a simple prescription for neutrino transport in their code they found that neutrino heating was insufficient to drive polar jets. A number of groups have studied the collapsar scenario using Newtonian MHD codes and implementing the Paczynski-Witta potential in order to approximate the gravitational field of central BH \\cite{PMAB03,FKYHS06,NTMT07}. In this approach it is impossible to capture the Blandford-Znajek effect and only the magnetic braking of the accretion disk can be investigated. The general conclusion of these studies is that the accretion disk can launch magnetically-driven jets provided the magnetic field in the progenitor core is sufficiently strong. Unfortunately, the jet power has not been given in most of these papers and is difficult to evaluate from the published numbers. In the simulations of Proga et al.\\shortcite{PMAB03} the jet power at $t\\simeq 0.25$s is $\\simeq 10^{50}\\mbox{erg}/$s. The initial magnetic field in these simulations is monopole with $B\\simeq 2\\times 10^{14}$G at $r=3r_g$, where $r_g=GM\\sub{bh}/c^2$ (private communication). \\begin{figure*} \\includegraphics[width=57mm]{figures/exp2.png} \\includegraphics[width=57mm]{figures/exp4.png} \\includegraphics[width=57mm]{figures/exp3.png} \\caption{Solution immediately before the explosion (t=0.24s). Left panel: the baryonic rest mass density, $log_{10}\\rho$, in $g/cm^3$ and the magnetic field lines; Middle panel: the ratio of gas and magnetic pressures, $log_{10}P/P_m$, and velocity direction vectors; Right panel: the ratio of azimuthal and poloidal magnetic field strengths, $log_{10}B^\\phi/B_p$, and the magnetic field lines. } \\label{f0} \\end{figure*} The study of collapsars in full GRMHD is still in its infancy. Sekiguchi \\& Shibata\\shortcite{SS07} studied the collapse of rotating stellar cores and formation of BH in the collapsar scenario. Their results show powerful explosions soon after the accretion disk is formed around the BH and the free falling plasma of polar regions collides with this disk. These explosions are driven by the heat generated as a result of such collision. However, the authors have not accounted for the neutrino cooling and the energy losses due to photo-dissociation of atomic nuclei. and the explosions could be similar in nature to the ``successful'' prompt explosions of early supernova simulations \\cite{bethe}. Mizuno et al.\\shortcite{MYKS04a,MYKS04b} carried out GRMHD simulations in the time-independent space-time of a central BH. The computational domain did not include the BH ergosphere and thus they could not study the role of the Blandford-Znajek effect~\\cite{K04a}. The energy losses have not been included and the equation of state (EOS) was a simple polytrope. These simulations were run for a rather short time, $\\simeq 280 r_g/c$ where $r_g=GM/c^2$, and jets were formed almost immediately due to unrealistically strong initial magnetic field. In this letter we describe the first results of axisymmetric GRMHD simulations of collapsars where we use realistic EOS~\\cite{TS00}, include the energy losses due to neutrino emission (assuming optically thin regime) and photo-dissociation of nuclei (see the details of micro-physics in Komissarov \\& Barkov 2007 ), use the computational domain that includes the BH horizon and its ergosphere, and run simulations for a relatively long physical time, up to 0.5s. The neutrino heating is not included. \\begin{figure*} \\includegraphics[width=57mm]{figures/rho1.png} \\includegraphics[width=57mm]{figures/rho2.png} \\includegraphics[width=57mm]{figures/rho3.png} \\caption{Solution on different scales at $t=0.45$s. The colour images show the baryonic rest mass density, $log_{10}\\rho$ in g/cm$^3$, the contours show the magnetic field lines, and the arrows show the velocity field.} \\label{f1} \\end{figure*} \\begin{figure*} \\includegraphics[width=57mm]{figures/b1.png} \\includegraphics[width=57mm]{figures/b2.png} \\includegraphics[width=57mm]{figures/b3.png} \\caption{The inner region at $t=0.45$s. Left panel: the magnetization parameter, $log_{10}P/P_m$, and the magnetic field lines; Middle panel: the ratio of azimuthal and poloidal magnetic field strengths, $log_{10}B^\\phi/B_p$, and the magnetic field lines; Right panel: the magnetic field strength, $log_{10}(B)$, and the magnetic field lines. } \\label{f2} \\end{figure*} ", "conclusions": "Our results provide strong support to the idea that magnetic fields can play a crucial role in driving powerful GRB jets and associated stellar explosions not only in the magnetar model but also in the collapsar model. The main energy source for the jets and explosions in our simulations is the rotational energy of black hole and it is released via the Blandford-Znajek mechanism. The measured rate of energy release, $\\dot{E} \\ge 2\\times10^{51}\\mbox{erg}\\,\\mbox{s}^{-1}$, can explain the energetics of even the shortest of long duration GRBs. The fact that the rotational energy of black hole, $E\\sub{bh}\\simeq\\mbox{few}\\times10^{53}\\mbox{erg}$, exceeds the typical explosion values derived from observations, $E\\simeq 10^{52}\\mbox{erg}$, suggests a self-regulating process in which the black hole activity ceases when the blast wave terminates further mass supply to the accretion disk. The full details of the simulations together with the results of parameter study will be presented elsewhere." }, "0710/0710.5418_arXiv.txt": { "abstract": "By applying recent results for the slab correlation time scale onto cosmic ray scattering theory, we compute cosmic ray parallel mean free paths within the quasilinear limit. By employing these results onto charged particle transport in the solar system, we demonstrate that much larger parallel mean free paths can be obtained in comparison to previous results. A comparison with solar wind observations is also presented to show that the new theoretical results are much closer to the observations than the previous results. ", "introduction": "Cosmic rays (CRs) interacting with turbulent magnetic fields get scattered and accelerated (see Melrose 1968, Schlickeiser 2002). The theoretical description of these scattering and acceleration processes are essential for understanding the penetration and modulation of low-energy cosmic rays in the heliosphere, the confinement and escape of galactic cosmic rays from the Galaxy, and the efficiency of diffusive shock acceleration mechanisms. A key factor in CR scattering are the properties of the magnetic fields. A standard approach is the assumption of a superposition of a mean magnetic field $\\vec{B}_0 = B_0 \\vec{e}_z$ and a turbulent component $\\delta \\vec{B} (\\vec{x})$. Whereas the mean field can easily be meassured in the solar system (here we find approximatelly $B_0 \\approx 4-5 nT$), the turbulent component has to be emulated by turbulence models. In the literature there is no consensus available about the true turbulence properties (see Cho \\& Lazarian 2005 for a review). In the solar system, however, some turbulence properties such as the wave spectrum can be obtained from meassurements (see e.g. Denskat \\& Neubauer 1983, Bruno \\& Carbone 2005). More unclear are the orientation of the turbulence wave vectors (also refered to as turbulence geometry) and the dynamical decorrelation of the magnetic fields. In a recent CR diffusion study (Shalchi et al. 2006) a slab/2D composite model was combined with a nonlinear anisotropic dynamical turbulence (NADT) model. This model can be used to reproduce meassured CR mean free paths parallel and perpendicular to the mean field $\\vec{B}_0$. The authors of this article assumed that the slab correlation time scale is independent of the wave vector $\\vec{k}$. In a recent study (Lazarian \\& Beresnyak 2006), however, it was shown that the slab time scale is indeed $\\vec{k}-$dependent.\\footnote{That study put to the test the idea of the damping of slab perturbations by the ambient turbulence in Yan \\& Lazarian (2002), Farmer \\& Goldreich (2004).} More precisely it was found that $t_c^{-1} = \\gamma_c = v_A \\sqrt{k_{\\parallel} / L}$; here we used the correlation time $t_c$, the correlation rate $\\gamma_c$, the Alfv\\'en speed $v_A$, and the outer scale of the turbulence $L$. It is the purpose of this article to apply this new result of the slab correlation time scale onto cosmic ray parallel diffusion. A comparison with solar wind observations of the parallel mean free path is also presented. It is demonstrated that we can find a much larger parallel mean free path if we employ the correlation time scale of Lazarian \\& Beresnyak (2006). In Section 2 we explain the turbulence model that is used in this article. In Section 3 a quasilinear description of cosmic ray scattering is combined with this turbulence model to derive analytic forms of the pitch-angle diffusion coefficient and the parallel mean free path. In Section 4 we evaluate these formulas numerically to compute diffusion coefficients and we also provide a comparison with previous results and solar wind observations. In the closing Section 5 our results are summerized. ", "conclusions": "The theoretical explanation of measured parallel mean free paths in the solar system is a fundamental problem of space science. In a recent article (Shalchi et al. 2006) it has been demonstrated the these observations can indeed be reproduced theoretically. By using recent results of turbulence theory (Lazarian \\& Beresnyak 2006) we further improved the dynamical correlation function which is a key input in transport theory considerations. It is demonstrated in this article that the improved slab correlation time scale (see Eq. (\\ref{Alexmodel})) leads to a much larger parallel mean free path (see Fig. \\ref{nadtf4}). This effect is important since it was argued in several previous articles that the theoretical parallel mean free path is too small (Palmer 1982, Bieber et al. 1994) in comparison with solar wind observations. Another problem of cosmic ray scattering theory is the importance of nonlinear effects. Whereas we have applied QLT in the current article it was argued in other papers (e.g. Shalchi et al. 2004) that nonlinear effects are important for parallel diffusion. However, these nonlinear effects are directly related to the interaction between charged particles and 2D modes. These 2D modes were neglected since we assumed pure slab fluctuations. Therefore, QLT can be applied and the results presented in this article should be valid. For non-slab models, where 2D modes are present, however, the applicability of QLT is questionable. It has to be subject of future work to explore the validity of QLT for realistic turbulence models such as dynamical turbulence models in non-slab geometry." }, "0710/0710.4522_arXiv.txt": { "abstract": "We directly measure the evolution of the intergalactic Lyman-$\\alpha$ effective optical depth, $\\tau_{\\rm eff}$, over the redshift range $2 \\leq z \\leq 4.2$ from a sample of 86 high-resolution, high-signal-to-noise quasar spectra obtained with Keck/ESI, Keck/HIRES, and Magellan/MIKE. We find that our estimates of the quasar continuum levels in the \\Lya~forest obtained by spline fitting are systematically biased low, but that this bias can be accounted for using mock spectra. The mean fractional error $\\langle \\Delta C/C_{\\rm true} \\rangle$ is $<1\\%$ at $z=2$, 4\\% at $z=3$, and 12\\% at $z=4$. We provide estimates of the level of absorption arising from metals in the \\Lya~forest based on both direct and statistical metal removal results in the literature, finding that this contribution is $\\approx6-9\\%$ at $z=3$ and decreases monotonically with redshift. The high precision of our measurement indicates significant departures from the best-fit power-law redshift evolution, particularly near $z=3.2$. ", "introduction": "The evolution of the intergalactic medium (IGM) as traced by the \\Lya~forest provides a powerful record of the thermal and radiative history of the Universe. This power owes to our ability to measure the \\Lya~opacity of the IGM as a function of redshift, as well as to the relatively simple physics of the \\Lya~forest. In fact, cosmological simulations in which the forest arises from absorption by smooth density fluctuations imposed on the warm photoionized IGM as a natural consequence of hierarchical structure formation within cold dark matter models \\citep[e.g.,][]{1996ApJ...457L..51H}, have been remarkably successful at reproducing the properties of the absorption observed in actual quasar spectra. This synergy between theory and observations make the \\Lya~forest a particularly compelling probe of the diffuse Universe.\\\\ \\\\ In this work, we present a direct precision measurement of the effective \\Lya~optical depth and its evolution over the redshift range $2\\leq z \\leq4.2$ from a sample of 86 high-resolution, high signal-to-noise quasar spectra obtained obtained with Keck/ESI (16), Keck/HIRES (44), and with Magellan/MIKE (26). The full details have been reported in \\citep{2007arXiv0709.2382F}. We assume a cosmology with $(\\Omega_{m},~\\Omega_{b},~\\Omega_{\\Lambda},~h,~\\sigma_{8})=(0.27,~0.046,~0.73,~0.7,~0.8)$ \\citep[][]{2007ApJS..170..377S}. ", "conclusions": "" }, "0710/0710.3591_arXiv.txt": { "abstract": "The near-infrared spectrum of (50000) Quaoar obtained at the Keck Observatory shows distinct absorption features of crystalline water ice, solid methane and ethane, and possibly other higher order hydrocarbons. Quaoar is only the fifth Kuiper belt object on which volatile ices have been detected. The small amount of methane on an otherwise water ice dominated surface suggests that Quaoar is a transition object between the dominant volatile-poor small Kuiper belt objects (KBOs) and the few volatile-rich large KBOs such as Pluto and Eris. ", "introduction": "While once Pluto and Triton were the only objects in the outer solar system known to contain volatile ices on their surfaces, the recent discoveries of frozen methane on the large Kuiper belt objects (KBOs) Eris, Sedna, and 2005 FY9 have shown that these objects are part of a new class of surface volatile rich bodies in the outer solar system \\citep{2005ApJ...635L..97B, 2006A&A...445L..35L,2005A&A...439L...1B, ebspec}. In contrast to these bodies with detectable volatiles, spectral observations of small KBOs over the past decade have found that most of these objects either contain varying amounts of involatile water ice on their surfaces or have flat spectra with no identifiable features \\citep{Kris}. To understand the dichotomy between volatile rich and volatile free surfaces in the outer solar system, \\citet{2007ApJ...659L..61S} constructed a simple model of atmospheric escape of volatile ices over the age of the solar system. They found that while most KBOs are too small and hot to retain their initial volatile ices to the present day, a small number are large and cold enough to retain these ices on their surfaces. As anticipated, the model suggests that the largest KBOs, Eris, Pluto, and Sedna are all expected to retain surface volatiles, while the vast majority of the other known objects in the Kuiper belt are expected to have lost all surface volatiles. Two known intermediate-sized KBOs are predicted to be in the transition region where they may have differentially lost some volatile ices (N$_2$) but retained others (CH$_4$). One of these transition objects, 2005 FY9, with a diameter of $\\sim$1450 km \\citep{2007astro.ph..2538S} does indeed appear to contain abundant CH$_4$ but be depleted in N$_2$. The other object that appears to be in the volatile non-volatile transition region is Quaoar, with a diameter of 1260$\\pm 190$ km\\citep{2004AJ....127.2413B}. The infrared spectrum of Quaoar does not resemble that of 2005 FY9, however. Quaoar's spectrum is dominated by absorptions due to involatile water ice, which is not detected at all on 2005 FY9. In addition, \\citet{2004Natur.432..731J} reported the detection of an absorption feature near 2.2 $\\mu m$ that they attributed to ammonia hydrate. They also detected the presence of crystalline water ice which, at the $\\sim$ 40 K radiative equilibrium temperature of Quaoar, is thought to be converted to amorphous water ice on a relatively short ($\\sim 10$ Myr) timescale by cosmic ray bombardment. The crystallinity of the water ice and the detection of the 2.2 $\\mu m$ feature that they attributed to ammonia hydrate led \\citet{2004Natur.432..731J} to suggest that Quaoar may have experienced relatively recent cryovolcanic activity. In this paper, we present a new infrared spectrum of Quaoar with a signal-to-noise in the K-band six times greater than that of \\citet{2004Natur.432..731J} and model the ices present on the surface. ", "conclusions": "With significantly higher signal-to-noise in the 2.0 - 2.4 $\\mu m$ region, the 2.2 $\\mu m$ absorption feature on Quaoar previously identified as ammonia hydrate \\citep{2004Natur.432..731J} is clearly seen to be due to methane ice. No compelling evidence is seen for the presence of ammonia. The presence of crystalline water ice on the surface of Quaoar still remains unexplained because it is expected that ice should currently exist in the amorphous form on the $\\sim$ 40 K surface of Quaoar. However, the presence of the 1.65 $\\mu m$ absorption feature due to crystalline water ice in the spectrum of every well observed water ice rich KBO (even down to diameters of only a few hundred kilometers) \\citep{Kris} suggests that exotic processes such as cryovolcanism are unlikely to be required. The presence of crystalline water ice on so many small outer solar system bodies may indicate that our current understanding of the physics of the crystalline/amorphous phase transition may not be complete. The spectrum of Quaoar is consistent with that of a cold geologically dead object slowly losing the last of its volatile ices by escape in a tenuous, perhaps patchy, atmosphere. Ethane is an expected by-product of irradiation of methane ice \\citep{2003NIMPB.209..283B}. The presence of ethane on Quaoar and on 2005 FY9 supports the suggestion of \\citet{ebspec} that these irradiation products are preferentially seen on bodies with large abundances of pure methane rather than on the bodies where the methane is diluted in nitrogen. Quaoar also appears to be rich in more complex irradiation products. Quaoar is the only water ice rich KBO which has a red color in the visible. Other water ice rich KBOs like Orcus, Charon, and 2003 EL61 and its family of collisional fragments are all blue in the visible \\citep{Kris}. Quaoar's red surface is likely due to the continued irradiation of methane, ethane, and their products on the surface \\citep{2006ApJ...644..646B}. While methane on Quaoar is sufficiently volatile that it is likely to seasonally migrate if Quaoar has a moderate obliquity, ethane and the other irradiation products are essentially involatile at Quaoar's temperature. Quaoar is therefore likely to have an irregular covering of irradiation products, perhaps leading to rotational variability in its visible color and in the abundance of ethane. Continued observations of this object will provide insight into the nature of the volatile non-volatile transition and atmospheric escape in the outer solar system. {\\it Acknowledgments:} We thank an anonymous referee for a helpful review. E.L.S. is supported by a NASA Graduate Student Research Fellowship. The data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, The University of California and the National Aeronautics and Space Administration. The observatory was made possible by the generous financial support of the W.M. Keck Foundation." }, "0710/0710.1846_arXiv.txt": { "abstract": "{The theory of stellar evolution can be more closely tested if we have the opportunity to measure new quantities. Nowadays, observations of galactic RR Lyr stars are available on a time baseline exceeding 100 years. Therefore, we can exploit the possibility of investigating period changes, continuing the pioneering work started by V.~P.~Tsesevich in 1969.} {We collected the available times of maximum brightness of the galactic RR Lyr stars in the GEOS RR Lyr database. Moreover, we also started new observational projects, including surveys with automated telescopes, to characterise the O--C diagrams better.} {The database we built has proved to be a very powerful tool for tracing the period variations through the ages. We analyzed 123 stars showing a clear O--C pattern (constant, parabolic or erratic) by means of different least--squares methods.} {Clear evidence of period increases or decreases at constant rates has been found, suggesting evolutionary effects. The median values are $\\beta$=+0.14~d~Myr$^{-1}$ for the 27 stars showing a period increase and $\\beta$=--0.20~d~Myr$^{-1}$ for the 21 stars showing a period decrease. The large number of RR Lyr stars showing a period decrease (i.e., blueward evolution) is a new and intriguing result. There is an excess of RR Lyr stars showing large, positive $\\beta$ values. Moreover, the observed $\\beta$ values are slightly larger than those predicted by theoretical models.} {} ", "introduction": "RR Lyr variables are low--mass stars in a core helium burning phase; they fill the part of the HR diagram where the horizontal branch intersects the classical instability strip. The crossing of the instability strip can take place in both directions; as a consequence, the periods will be either increasing, if the stars evolve from blue to red, or decreasing, if they evolve from red to blue. Despite its importance as a test for the stellar evolution theory, the {\\it observed} rate of the period changes is still an unknown quantity. It has been measured in globular clusters (e.g., Smith \\& Sandage \\cite{mfifteen}, Lee \\cite{lee}, Jurcsik et al. \\cite{omega}), comparing periods determined in different epochs. However, this task has not been undertaken yet on the wide population of galactic RR Lyr stars, though most of them have been studied for tens of years, several of them since the end of the XIX$^{\\rm th}$ century, and therefore the available data are almost continuous. One of the critical points is to determine the importance of the negative rates, i.e., the period decreases. Such rates have been observed in some cluster stars (Jurcsik et al. \\cite{omega}), but it is unclear in which evolutionary phase they occur (Sweigart \\& Renzini \\cite{random2}). In the present paper, we derive the period variation rate of the best observed RR Lyr stars belonging to the RRab sub-class. For this purpose we use the {\\it GEOS RR Lyr database} which is described in Sect.~\\ref{sect_grrdb}. The observations of the TAROT telescopes coordinated in the {\\it GEOS RR Lyr Survey} give a strong impulse to this analysis, providing a large number of times of maximum brightness in the recent years. They complete the effort of the amateur observers, in particular in the European associations BAV (Bundesdeutsche Arbeitsgemeinschaft f\\\"ur Ver\\\"anderliche Sterne) and GEOS (Groupe Europ\\'een d'Observations Stellaires), which have been surveying these stars for years. These new observations are described in Sect.~\\ref{sect_grrs}. Next, Sect.~\\ref{sect_datan} shows how the data were analysed. In Sect.~\\ref{sect_cstrate}, we analyse the stars with constant period variation. Section~\\ref{sect_BLZH} is devoted to the study of some particularities encountered during the present work: Blazhko and light--time effects. Finally, a discussion of the results is developed in Sect.~\\ref{sect_discuss} in the light of stellar evolution. ", "conclusions": "The analysis of O--C variations over a timescale of 100 years or more has proved to be a powerful tool for providing quantitative tests of the stellar evolution of the horizontal branch stars. We can stress some well--established observational facts: \\begin{enumerate} \\item RRab stars experiencing blueward evolution (i.e., period decreases) are quite common, only slightly less so than RRab stars experiencing redward evolution (i.e., period increases). The period ranges covered by the two groups are very similar and the mean and median period values are nearly coincident; \\item the absolute values of period changes are larger than expected. Also in the extreme case of the rapid--evolving star \\object{SV Eri} the rate is much larger than expected; \\item the O--C behavior can be very complicated in some cases, with abrupt or irregular period variations, rather than monotonic. The regular variations caused by the light--time effect (and hence duplicity), often invoked to explain large O--C excursions, are not convincingly observed in our extensive sample. The physical explanations should be searched in the stellar structure; \\item Blazhko effect is often superimposed on secular changes, but the monotonic trend due to evolutionary variations still remains visible. \\end{enumerate} As a general conclusion about the comparison between our observational results and the theoretical predictions, we claim that there is a very powerful feedback between the two approaches. In particular, theoretical investigations should take into account that we have observational evidence of many RR Lyr stars showing blueward evolution. The theoretical models should also match the observed $\\beta$ values in a more satisfactory way, as these seem to be higher than expected, both for redward and blueward evolutions. On the other hand, the observational effort to monitor RR Lyr stars should be continued, possibly extended to stars at fainter magnitudes, and the accuracy of the maximum time determinations should be improved (by using automated telescopes), thus obtaining the same information on a shorter time baseline." }, "0710/0710.1077_arXiv.txt": { "abstract": "In this work we give special attention to the bimetric theory of gravitation with massive gravitons proposed by Visser in 1998. In his theory, a prior background metric is necessary to take in account the massive term. Although in the great part of the astrophysical studies the Minkowski metric is the best choice to the background metric, it is not possible to consider this metric in cosmology. In order to keep the Minkowski metric as background in this case, we suggest an interpretation of the energy-momentum conservation in Visser's theory, which is in accordance with the equivalence principle and recovers naturally the special relativity in the absence of gravitational sources. Although we do not present a general proof of our hypothesis we show its validity in the simple case of a plane and dust-dominated universe, in which the `massive term' appears like an extra contribution for the energy density. ", "introduction": "\\label{intro} Could the graviton have a non-zero rest mass? The observations have shown that this is a possibility. One of the most accurate bounding on the mass of the graviton comes from the observations of the planetary motion in the solar system. Variations on the third Kepler law comparing the orbits of Earth and Mars can lead us to $m_g < 7.8 \\times 10^{-55}g$ \\cite{Talmadge88}. Another bound comes from the analysis of galaxy clusters that lead to $m_g < 2 \\times 10^{-62}g$ \\cite{Goldhaber74} which is considerably more restrictive but less robust due to the uncertainties in the content of the universe in large scales. Studying rotation curves of galactic disks, \\cite{JC} has found that we should have a massive graviton of $m_g \\ll 10^{-59}g$ in order to obtain a galactic disk with a scale length of $b\\backsim10$ kpc. The above tests are obtained from static fields based on deviations of the newtonian gravity. In the weak field limit has been proposed \\cite{Finn2002} to constraint $m_g$ using data on the orbital decay of binary pulsars. From the binary pulsar PSR B1913+16 (Hulse-Taylor pulsar) and PSR B1534+12 it is found the limit $m_g < 1.4 \\times 10^{-52}g$, which is weaker than the bounds in static field. It is worth recalling that the mass term introduced via a Pauli-Fierz (PF) term in the linearized approximation produces a theory whose predictions do not reduce to those of general relativity for $m_g \\rightarrow 0$. This is the so called van Dam Veltmann Zakharov discontinuity \\cite{Veltman1970}. Moreover the Minkowski space as background metric is unstable for the PF theory \\cite{Gruzinov2005}. However, there is no reason to prefer the PF term over any other non-PF quadratic terms. It is important to emphasize that these mass terms do not have clear extrapolation to strong fields. A way to do that was proposed by Visser \\cite{visser98}. To generalize the theory to strong fields, Visser makes use of two metrics, the dynamical metric ($g_{\\mu\\nu}$) and a non-dynamical background metric ($\\left( g_0\\right) _{\\mu\\nu}$) that are connected by the mass term. Although adding a prior geometry is not in accordance with the usual foundations underlying Einstein gravity, it keeps intact the principles of equivalence (at least in its weak form) and general covariance in the Visser's work. Some interesting physical features emerge from the theory such as extra states of polarizations of the gravitational waves \\cite{wayne2004}. In the present article, we explore some aspects which are not treated by Visser in his original paper. In the great part of the astrophysical studies the Minkowski metric is the most appropriate choice to the background metric. However, in the study of cosmology, it is not possible to consider this kind of metric, and we need some prior considerations regarding a background metric. Once this problem emerges from the coupling of the two metrics and the energy's conservation condition, we analyze an alternative interpretation of this condition. We also show that this interpretation is in accordance with the equivalence principle and recovers naturally the special relativity in the absence of gravitational sources. Arguments in favor of a Minkowskian background metric in Visser's theory are also considered. This paper is organized as follows: in section \\ref{sec:1} we show how to introduce a mass for the graviton through a non-PF term. We present the strong field extrapolation as given by Visser in section \\ref{sec:2}. In section \\ref{sec:3} we show that the theory is not in accordance with a Minkowski background metric in the study of cosmology. In section \\ref{sec:4} we re-interpret the stress-energy conservation in order to keep Minkowski as background in any case. In particular, we show that our re-interpretation is in accordance with the equivalence principle. In section \\ref{sec:5} we show why Minkowski is the most natural choice to the background metric. We briefly study some cosmological consequences of our interpretation of the energy-momentum conservation in section \\ref{sec:6}. And finally, we present our conclusions in the last section. ", "conclusions": "\\label{sec:7} Our interpretation of the energy-momentum conservation in the Visser's massive gravity is in accordance to the equivalence principle and recover naturally the results of special relativity in the absence of gravitational sources. The point of view considered in this paper allow us to consider Minkowski as background metric in Visser's theory in all astrophysical cases including cosmology. This new interpretation may lead to interesting cosmological results once we can construct a cosmological model in a theory with massive gravitons with a Minkowski background. Additional contributions to the cosmological fluids will appear due to the modifications in the interaction potential, which, maybe, would be a way of treat the dark-energy problem. The analyses of the theory in the absence of gravitational sources lead us to exclude the de Sitter space-time as a vacuum solution of the massive gravity, once a constant $\\Lambda$ term is rigorously zero in a flat background. Another interesting feature is that our interpretation of the energy conservation in strong fields is independent of the form of the tensor which interact with the perfect-fluid tensor, so this can be used to other models with additional energy-momentum contribution." }, "0710/0710.0247_arXiv.txt": { "abstract": "We present high spatial resolution spectroscopic measurements of dynamic fibrils (DFs) in the Ca~{\\small{II}}~8662~{\\AA} line. These data show clear Doppler shifts in the identified DFs, which demonstrates that at least a subset of DFs are actual mass motions in the chromosphere. A statistical analysis of 26 DFs reveals a strong and statistically significant correlation between the maximal velocity and the deceleration. The range of the velocities and the decelerations are substantially lower, about a factor two, in our spectroscopic observations compared to the earlier results based on proper motion in narrow band images. There are fundamental differences in the different observational methods; when DFs are observed spectroscopically the measured Doppler shifts are a result of the atmospheric velocity, weighted with the response function to velocity over an extended height. When the proper motion of DFs is observed in narrow band images, the movement of the top of the DF is observed. This point is sharply defined because of the high contrast between the DF and the surroundings. The observational differences between the two methods are examined by several numerical experiments using both numerical simulations and a time series of narrow band H$\\alpha$ images. With basis in the simulations we conclude that the lower maximal velocity is explained by the low formation height of the Ca~IR~line. We conclude that the present observations support the earlier result that DFs are driven by magneto-acoustic shocks exited by convective flows and p-modes. ", "introduction": "The dynamical nature of the chromosphere is obvious when the Sun is imaged in the line center of strong chromospheric spectral lines, most commonly in the H$\\alpha$ line \\citep[e.g.,][]{2006Noort}. One of the dominating features is a vast number of thin (0.2-1 Mm) omnipresent jet like structures \\citep{1968Beckers}. On the quiet solar limb they are commonly known as spicules, while on the quiet disk they are often called mottles, and finally in active regions they are known as active region fibrils or dynamic fibrils (DFs). The nomenclature can be confusing, but there are strong indications that these structures are physically closely related \\citep{1994Tsiropoula,1995Suematsu,2001Christo,2007Rouppe}. Recently, a combination of high--resolution observations and advanced numerical modeling have shown that DFs are most likely driven by shocks that form when photospheric oscillations leak into the chromosphere along inclined flux tubes \\citep{1990Suematsu,2004dePontieu,2006Hansteen,2007dePontieu}. The inclination of the magnetic field lowers the acoustic cutoff frequency sufficiently to allow p--modes with the dominant low frequencies to propagate along flux tubes \\citep{1973Mich,1977Bel}. These insights into the formation of DFs have become possible because of recent developments in observational techniques, such as bigger telescopes combined with real time wavefront corrections by adaptive optics (AO) systems \\citep[e.g.][]{2000Rimmele,SSTAO}, and post-processing methods \\citep[e.g.][]{speckle,momfbd} which have made observations of these jet structures much more reliable. These developments have spurred several authors to focus on the detailed understanding of DFs \\citep[e.g.,][]{2003dePontieu,2004dePontieu, 2005dePontieu,2007dePontieu,2004Kostas,2006Hansteen,2006deWijn,2007Julius, 2007Lars}. One of the important results of this work is that the DFs are driven by and can channel photospheric oscillations into the chromosphere and the corona. \\begin{figure*}[!ht] \\includegraphics[width=\\textwidth]{f1.eps} \\caption{ Spectrograms of the H$\\alpha$ line (panel a) and the Ca~{\\small{II}}~8662~{\\AA} line (panel b). Notice the highly dynamical line center of the Ca line. The corresponding MFBD processed slitjaw image (panel c) and the narrow band H$\\alpha$ DOT image (panel d). The position of the spectrograph's slit is marked with a line in the DOT image. } \\label{plotone} \\end{figure*} In two papers \\citet{2006Hansteen} and \\citet{2007dePontieu} used high spatial and high cadence observations of the H$\\alpha$ line center together with realistic simulations to investigate the nature of DFs. One of their conclusions was that the DFs follow parabolic paths along their axis with decelerations lower than the solar gravitational deceleration. Previous observations did not allow an accurate determination of the nature of the trajectory because of lower quality data and line--of--sight (LOS) effects that were difficult to estimate \\citep{1988Nishikawa,1995Suematsu}. \\citet{2006Hansteen} and \\citet{2007dePontieu} further report regional differences between DFs observed in two different plage areas. The regional differences are explained by different inclination angles of the magnetic fields in the two regions. In this way the magnetic topology of the solar atmosphere works as a filter, where only waves with certain periods can leak through. The simulations, spanning from the upper convection zone to the corona, reproduce the observed correlations between the maximum velocities and decelerations in DFs leading to the conclusion that DFs are formed by chromospheric shocks driven by global p-modes and convective flows. In this paper we add to the understanding of the DFs, by analyzing high spatial and high cadence spectrograms of the Ca {\\small{II}} 8662 {\\AA} line, put into context by simultaneous H$\\alpha$ spectrograms and narrow-band images. Furthermore, the observational results are compared with the numerical simulations of \\citet{2006Hansteen}. In \\S~\\ref{Obs} we describe the observing program and instrumentation. The data reduction method is described in \\S~\\ref{Data}. In \\S~\\ref{Obsres} we show the results of the observations. The main observational errors are discussed in \\S~\\ref{error}. To get a better understanding of the observations we present several numerical experiments in \\S~\\ref{Sim}. Finally, we summarize the results in \\S~\\ref{Con}. ", "conclusions": "\\label{Con} In this work we have presented co-spatial and co-temporal narrow band H$\\alpha$ images and spectroscopic measurements of the Ca~{\\small{II}}~8662~{\\AA} line. These observations have been used to identify 26 DFs, and measure their Doppler shifts. A reduced--$\\chi^2$ analysis shows that the time evolution of the Doppler shifts are well approximated by a linear fit, if the measurement errors are about $2$~km\\,s${}^{-1}$. Using this approximation we derive values for the decelerations and maximal velocities for each DF. Scatter plots of the deceleration and maximal velocity show a strong positive correlation between the two. We also observe weak correlations between the deceleration and lifetime and the maximal velocity and lifetime. These results are supporting the shock-wave theory as explanation model for the DFs \\citep{2006Hansteen, 2007dePontieu,2007Lars}. Furthermore, the Doppler shifts show that at least a subset of DFs are caused by mass moving up and down in the atmosphere. The values of the maximum velocity and decelerations are all somewhat lower than earlier reported values. Using numerical experiments we have explained the differences in the two observational sets with the intrinsic differences in observational methods. Earlier observations have used the high contrast seen between the top of the DF and the background for measuring the proper motion of the DF. This high contrast is caused by the intensity increase due to contributions from the transition region. In the present observations the DF motion is measured using Doppler shifts, which are affected by the atmospheric conditions over the formation height. The formation height of the Ca~{\\small{II}}~8662~{\\AA}~line is much lower than the transition region. Since the shock amplitude is increasing from the formation height for the Ca IR line to the transition region we necessarily measure lower velocities using spectroscopy. The difference in maximal velocities derived from the two methods in our simulations is about a factor two, which is about the same as observed. The one dimensional nature of the slit spectrograph somewhat affects our results, but experiments with narrow band images show that this is not altering the results significantly." }, "0710/0710.5712_arXiv.txt": { "abstract": "The cosmological concordance model contains two separate constituents which interact only gravitationally with themselves and everything else, the dark matter and the dark energy. In the standard dark energy models, the dark matter makes up some 20\\% of the total energy budget today, while the dark energy is responsible for about 75\\%. Here we show that these numbers are only robust for specific dark energy models and that in general we cannot measure the abundance of the dark constituents separately without making strong assumptions. ", "introduction": "We cosmologists are very proud that our field has finally reached the status of ``precision science'' over the last decade. The quality of the current observations of the cosmic microwave background (CMB), the galaxy distribution and the luminosity distance to type Ia supernovae (SN-Ia) is indeed impressive, and has allowed the construction of a concordance model in which the universe contains the known, baryonic, matter (5\\% of the energy density today), radiation (negligible energy density today), dark matter (20\\%) and dark energy (75\\%). The need for dark matter became apparent long ago in order to explain the motion of galaxies in clusters \\cite{darkmat} and the observed galaxy rotation curves. In cosmology, it is often modelled as a pressureless fluid with negligible interactions. Much more recently, less than ten years ago, new SN-Ia data \\cite{sn1a} convinced the majority of cosmologists that dark energy was needed as well. Until today the nature of the dark energy is a deep mystery. Although many models have been proposed, there are none that can explain its current abundance in a natural way. The alternative to the model building is a more phenomenological approach, where one measures the physical properties of the dark energy. To this end, one introduces a completely general fluid and tries to determine its characteristics from observations. ", "conclusions": "We have seen that cosmology cannot measure separately the properties of the dark matter and of a general dark energy component. In order to do that, we either need to impose additional assumptions, for example that the dark energy is a scalar field, or else we need a non-gravitational measurement of the dark matter properties, specifically of its contribution to the total energy density of the universe. One possibility is a detection of supersymmetry at LHC, which may in turn determine the abundance and mass of the lightest stable SUSY particle, one of the best candidates for the dark matter. As a corollary, if the abundance determined in this way is not the one expected within the $\\Lambda$CDM cosmological concordance model, one possible explanation is an evolving dark energy. We can also consider the degeneracy as a test of the generality of the different approaches to measure the dark energy equation of state. Since no analysis so far seems to have found it, we can only wonder what else has been overlooked. Finally, although we show here that one never can prove experimentally from cosmological data alone that the dark energy is a cosmological constant, it is remarkable that a model containing just cold dark matter and $\\Lambda$ fits the data so well. From a model selection point of view $\\Lambda$CDM is still the preferred model because of its simplicity. \\ack It is a pleasure to thank Luca Amendola, Ruth Durrer, Domenico Sapone and Anze Slosar for interesting discussions. MK acknowledges funding by the Swiss NSF." }, "0710/0710.0137_arXiv.txt": { "abstract": "We present multiple epochs of H$\\alpha$ spectroscopy for 47 members of the open cluster NGC 3766 to investigate the long term variability of its Be stars. Sixteen of the stars in this sample are Be stars, including one new discovery. Of these, we observe an unprecedented 11 Be stars that undergo disk appearances and/or near disappearances in our H$\\alpha$ spectra, making this the most variable population of Be stars known to date. NGC 3766 is therefore an excellent location to study the formation mechanism of Be star disks. From blue optical spectra of 38 cluster members and existing Str\\\"omgren photometry of the cluster, we also measure rotational velocities, effective temperatures, and polar surface gravities to investigate the physical and evolutionary factors that may contribute to the Be phenomenon. Our analysis also provides improvements to the reddening and distance of NGC 3766, and we find $E(B-V) = 0.22 \\pm 0.03$ and $(V-M_{\\rm V})_0 = 11.6 \\pm 0.2$, respectively. The Be stars are not associated with a particular stage of main-sequence evolution, but they are a population of rapidly rotating stars with a velocity distribution generally consistent with rotation at $70-80$\\% of the critical velocity, although systematic effects probably underestimate the true rotational velocities so that the rotation is much closer to critical. Our measurements of the changing disk sizes are consistent with the idea that transitory, nonradial pulsations contribute to the formation of these highly variable disks. ", "introduction": "\\setcounter{footnote}{3} NGC 3766 is a rich, young open cluster in the Carina spiral arm that is well known for its high content of Be stars \\citep{slettebak1985}, and many previous studies of this cluster have focused on the characteristics of these stars to identify their evolutionary status. The cluster has been the target of numerous photometric studies \\citep{ahmed1962, yilmaz1976, shobbrook1985, shobbrook1987, moitinho1997, piatti1998, tadross2001, mcswain2005b}. But despite these intensive investigations, the cluster's age and distance remain somewhat uncertain; measurements of its age range from 14.5 to 25 Myr (WEBDA\\footnote{The WEBDA database is maintained by E.\\ Paunzen and is available online at http://www.univie.ac.at/webda/navigation.html.}; \\citealt{lynga1987, moitinho1997, tadross2001}), and its distance is between 1.5 and 2.2 kpc. The reddening $E$($B$-$V$) is between 0.16 and 0.22 (see the discussion of \\citealt{moitinho1997}). Spectroscopic investigations of NGC 3766 have targeted a limited sample of cluster members, focusing primarily on the Be star and supergiant populations (\\citealt{harris1976}; \\citealt{mermilliod1982} and references therein; \\citealt{slettebak1985, levesque2005}). Even the eclipsing double-lined spectroscopic binary BF Centauri (= HD 100915), a member of NGC 3766, has been largely neglected by modern spectroscopic observations (\\citealt{clausen2007} and references therein). For most cluster members, no detailed information about their physical characteristics such as temperature, gravity, rotation, and metallicity are known. In this work, we present red and blue optical spectra for both normal B-type and Be stars in the cluster. Like many prior studies of NGC 3766, our primary goal is to investigate the Be star population; but unlike other works, we achieve a more complete understanding of this subset of B stars by comparing these emission-line objects to their non-emission counterparts. Therefore we present measurements of the effective temperature, $T_{\\rm eff}$, surface gravity, $\\log g$, and in most cases the projected rotational velocity, $V \\sin i$, for 26 normal B stars and 16 Be stars in NGC 3766. We use these results to improve the known reddening and distance to the cluster. From multiple epochs of H$\\alpha$ spectroscopy, we also investigate the variability of the circumstellar disks and estimate the disk mass loss/gain rates for 11 Be stars. Finally, we use the observed disk masses and angular momenta to show that nonradial pulsations are a possible origin for the disks, and they probably fill during short-lived bursts of mass flow from the stellar surface. ", "conclusions": "Our spectroscopic analysis of NGC 3766 has revealed that Be stars may be much more common than we originally thought. In our photometric study of NGC 3766 \\citep{mcswain2005b}, we found up to 13 Be stars (5 definite, 8 uncertain) out of an expected 191 B-type stars, not counting the one Be star that saturated our photometry. The new total of 16 Be stars is 23\\% greater. Among these 16 Be stars, 2--5 of them appear to have almost no disk at any given time, and an additional 2--4 have extremely subtle emission in their H$\\alpha$ line profile that could easily be mistaken for other phenomena (such as NRP manifesting themselves as bumps moving across the line or SB2 line blending). Therefore 25--50\\% of the Be stars may go undetected in a single spectroscopic observation, and photometric snapshots are even less likely to discern such weak emitters. We note four stars (Nos.\\ 27, 45, 49, and 77) that were found to be possible or likely Be stars in the photometric study by \\citet{shobbrook1985, shobbrook1987}, but they never showed emission during our observations and thus remain unconfirmed. The existence of transitory, weak disks (especially Nos.\\ 130 and 196) could mean that many more Be stars are waiting to be discovered. For our total sample of 48 Southern open clusters in our photometric survey, we found a low Be fraction of $2-7$\\% \\citep{mcswain2005b}. Considering the very weak disks that are observed in NGC 3766 and the exceptionally high variability among the cluster's Be population, the total fraction of Be stars could be much greater. We are currently performing a similar spectroscopic study of several other clusters from our survey, and we will address those results in a future paper. While the Be stars of NGC 3766 are not distinguishable from normal B-type stars by their evolutionary states, they do form a population of rapidly rotating stars. With two exceptions, their measured velocities are consistent with a uniform population of rapid rotators having $V = 0.7-0.8 \\; V_{\\rm crit}$. Gravitational darkening and weak emission in the \\ion{He}{1} lines may mean that these velocities are underestimated by as much as 33\\% \\citep{townsend2004}, so the true $V_{\\rm rot}$ is probably at least $0.84 \\; V_{\\rm crit}$. From the measured changes in the disks' masses and angular momenta, NRP are a capable source for the mass flow into the equatorial plane. The pulsations may be a transitory phenomenon, however, and the variable nature of the Be stars probably reflects dramatic changes in the surface activity." }, "0710/0710.2903_arXiv.txt": { "abstract": "We use the kinetic theory of nucleation to explore the properties of dust nucleation in sub-saturated vapors. Due to radiation losses, the sub-critical clusters have a smaller temperature compared to their vapor. This alters the dynamical balance between attachment and detachment of monomers, allowing for stable nucleation of grains in vapors that are sub-saturated for their temperature. We find this effect particularly important at low densities and in the absence of a strong background radiation field. We find new conditions for stable nucleation in the $n-T$ phase diagram. The nucleation in the non-LTE regions is likely to be at much slower rate than in the super-saturated vapors. We evaluate the nucleation rate, warning the reader that it does depend on poorly substantiated properties of the macro-molecules assumed in the computation. On the other hand, the conditions for nucleation depend only on the properties of the large stable grains and are more robust. We finally point out that this mechanism may be relevant in the early universe as an initial dust pollution mechanism, since once the interstellar medium is polluted with dust, mantle growth is likely to be dominant over non-LTE nucleation in the diffuse medium. ", "introduction": "Dust particles are one of the fundamental components of the interstellar medium (ISM) and an ever-present worry for observers due to their opacity at optical and UV wavelengths (Cardelli, Clayton \\& Mathis 1989). The ISM of the Milky Way is polluted by a mixture of grains made of a variety of materials, likely dominated by carbonaceous grains, silicates, and small PAHs particles (Mathis, Rumpl \\& Nordsieck 1977; Weingartner \\& Draine 2001). The dust properties are supposed to be the result of dust formation in the outflows of evolved stars (e.g. Salpeter1977; Stein \\& Soifer 1983; Mathis 1990; Whittet 1992; Draine 2003, and references therein) and subsequent evolution, and eventual dissolution, in the ISM, mainly as the effect of shock waves that destroy the grains through sputtering (Draine 1989; McKee 1989; Edmunds 2001). Alternative dust production sites are supernova explosions (Kozasa, Hasegawa \\& Nomoto 1989, 1991; Todini \\& Ferrara 2001; Nozawa et al. 2003; Schneider, Ferrara \\& Salvaterra 2004; Bianchi \\& Schneider 2007) and quasar outflows (Elvis, Marengo \\& Karovska, 2002). The theory of dust nucleation in astrophysics is heavily influenced by the theory of the nucleation of phase transitions in super-saturated vapors (Becker \\& Doring 1935; Feder et al. 1966; Abraham 1974). The theory had mild success in reproducing nucleation rates, but is still controversial in many aspects, especially because it extrapolates the properties of macroscopic bodies to clusters of few molecules and because it extends the thermodynamic approach to systems with a handful of particles. To add to these problems, astrophysical dust nucleation requires chemical reactions, since grains of materials that do not have a vapor state do nucleate (think, for example, to the nucleation of olivines from silicon oxides and metals; Draine 1979; Gail \\& Sedlmayr 1986). An alternative approach is the so-called kinetic theory, which describes nucleation as the result of attachment and detachment of monomers from a seed cluster of $n$ particles (atoms, molecules or radicals; Nowakowski \\& Ruckenstein 1991ab). Both the thermodynamic and the kinetic nucleation theory have been developed in conditions of {\\it true equilibrium}, i.e., when the two phases have the same temperature. In the astrophysical scenario, however, the temperature of the dust grains can be sensibly lower than the temperature of the gas in which they are embedded due to efficient radiation cooling (e.g., Draine 1981). This would seem to be irrelevant to nucleation theory, since a vapor needs to be already nucleated in order to have grains that can be colder than the gas phase. Even a sub-saturated vapor, however, has a large number of unstable clusters that form by random association of monomers (and rapidly evaporate). In this paper we study the effect of cooling of these proto-clusters in a sub-saturated vapor, and the effect this has on the balance between attachment and detachment of monomers. Using the kinetic theory of nucleation, we find that even largely sub-saturated vapors can nucleate, provided they are not immersed in a strong radiation field. We compute, albeit under some controversial assumptions, the non-LTE nucleation rate. We show that, even though it is not as large as in super-saturated vapors, it can produce dust grains at a rate that can reproduce the average dust grain density in the Milky Way over a timescale of several million years. In addition we show that non-LTE effects can increase the nucleation rate in super-saturated vapors. Non-LTE nucleation could therefore provide a slow channel for dust formation, in which dust is built over a relative long time in a slowly evolving region. Such an example could be the outflow from AGN nuclei (Elvis et al. 2002). Such evolution is different from the one envisaged in the classical dust factories -- AGB star atmospheres and supernov\\ae\\ -- where dust nucleation is rapid but short lived since the favorable conditions are rapidly lost. This paper is organized as follows: in \\S~2 we briefly review the classical kinetic theory of nucleation; in \\S~3 we compute the dust grain temperature and in \\S~4 we compute the new conditions for nucleation. In \\S~5 we consider the nucleation rate and discuss our results in \\S~6. ", "conclusions": "We have considered the effect of grain (cluster, droplet) cooling in the nucleation of liquid and solid phases in vapors. We find that the effect can be dramatic on the nucleation rate and on the nucleation phase diagram, allowing for nucleation in large regions of the parameter space that are classically considered to be non-nucleating. As is in general true for nucleation, there are several limits and approximations that we should bear in mind when considering the theory from the quantitative point of view. As exemplified by the comparison of the data with the theory in Fig.~\\ref{fig:j}, several orders of magnitude can separate the nucleation rate prediction from the observations. The controversial points are: \\begin{itemize} \\item {\\it Sticking coefficients} --- Most of the figures and computations in this work assume $k_s=1$. This is not always true (Batista et al. 2005). In addition to its dependence on temperature, the sticking coefficient may depend on the size of the cluster. A big cluster could more easily absorb the extra kinetic energy of the incoming monomer, compared to a small cluster (K00), and therefore $k_s$ may be significantly smaller than unity for very small clusters. \\item {\\it Capillary approximation} --- The capillary approximation, i.e., the assumption that the surface tension does not depend on the cluster size, is very controversial, and a change in the surface energy for very small clusters could result in big changes on the nucleation rate. For example, the discrepancy in Fig.~\\ref{fig:j} could be solved by assuming a larger surface tension for the very small water droplet. In addition, macromolecules do not even have a properly defined surface, and the whole concept does not apply. Finally, the exponential factor in Eq.~\\ref{eq:sigma} depends on the assumption that the clusters are spherical. A different ratio of the surface to the volume would modify this term. This is likely for small graphite clusters, since graphite tends to aggregate in a planar form. \\item {\\it Detachment rate for small clusters} --- In this paper, and in most nucleation theory, the detachment rate is computed by propagating to very small clusters the detachment rate of macroscopic bodies. It is very likely that, in the very small limit, the detachment is governed by completely different processes. Let us analyze the two limits. For a macroscopic body, the number of monomers is so large that monomers with a statistically higher energy can detach since their energy is larger than the binding energy. In the opposite limit of a dimer, the detachment has to be due to an external action: either a collision with a fast monomer or with a photon or with another cluster. In this limit destructive collisions have to be considered. \\item {\\it Cooling of macromolecules} --- This is a new problem that arises when the cooling of the grains is considered. We have assumed that the grains cool as modified black bodies down to the smallest sizes. When the number of monomers in the cluster becomes small, the cooling will not be through a continuum spectrum, but through lines and bands. A more refined treatment of cooling is necessary to compute accurate nucleation rates. \\item It should finally be kept in mind that we allowed for the complete cooling of the grains, neglecting the effects of a background radiation field in setting a lower limit to the temperature. In addition, we neglected the fact that the temperature of small grains is largely stochastic and that at high temperature some collisions between grains and gas particles can result in sputtering rather than accretion. \\end{itemize} Despite all these caveats, the main result of this paper holds: the region where a vapor spontaneously nucleate is not limited to the classic region, when nucleation takes place in thermal equilibrium and $T_g=T_s$. Allowing for the clusters to cool inhibits the detachment of monomers from the clusters and allow for nucleation at higher temperatures and lower densities than in the classical scenario. The nucleation rates tend to be orders of magnitude smaller than those derived in thermal equilibrium, but provide a non-negligible correction to the LTE rate for mildly super-saturated vapors with $S\\gsim1$, especially at low temperature. In the astrophysical scenario small nucleation rates are not a big worry. The ISM of our Galaxy contains approximately 1 per cent of its mass in dust grains (Mathis et al. 1977). This corresponds to approximately one dust particle every cubic meter (assuming a power-law grain size distribution as in Mathis et al. 1977). However, in the present day Universe the ISM is already polluted with dust and in the presence of seed grains it is likely that the process of mantle growth dominates over non-LTE nucleation in the diffuse medium. The process of non-LTE nucleation may therefore be important at high redshift, when the ISM is first polluted with metals by supernova explosions. It is unclear if supernov\\ae\\ do generate dust by themselves, and even more whether the generated dust can survive the sputtering in the forward-reverse shock systems (Bianchi \\& Schneider 2007; Nath, Laskar \\& Shull 2007). In the case that supernov\\ae\\ do mainly pollute the ISM with metals but with no or a negligible quantity of dust, non-LTE nucleation could become the dominant process of dust nucleation in the early universe. Detailed estimates of nucleation in the various scenarios require a more detailed understanding of the properties of the very small nuclei and are beyond the scope of this paper." }, "0710/0710.2418_arXiv.txt": { "abstract": "{ We report here on a new VHE source, HESS~J1908+063, disovered during the extended H.E.S.S. survey of the Galactic plane and which coincides with the recently reported MILAGRO unidentified source MGRO~J1908+06. The position, extension and spectrum measurements of the HESS source are presented and compared to those of MGRO~J1908+06. Possible counterparts at other wavelenghts are discussed. For the first time one of the low-lattitude MILAGRO sources is confirmed. } \\begin{document} ", "introduction": "H.E.S.S. observations of the inner Galactic plane in the [$270^{\\circ}$, $30^{\\circ}$] longitude range have revealed more than two dozens of new VHE sources, consisting of shell-type SNRs, pulsar wind nebulae, X-ray binary systems, a putative young star cluster, etc, and yet unidentified objects (see e.g. \\cite{HESSScanII} and \\cite{HESSSurveyICRC07} in these proceedings for a summary). The extended H.E.S.S. survey in the [$30^{\\circ}$-$60^{\\circ}$] longitude range performed between 2005 and 2007 overlaps with regions covered by the MILAGRO sky survey at longitudes greater than $30^{\\circ}$. The latter experiment has recently reported \\cite{MILAGRO} three low-latitude sources including, MGRO~J1908+06, detected after seven years of operation (2358 days of data) at 8.3$\\sigma$ (pre-trials) confidence level. MGRO~J1908+06, of which the extension remains unknown but bounded to a maximum diameter of 2.6$^{\\circ}$, is located near the galactic longitude $\\sim 40^{\\circ}$ and hence is covered by the H.E.S.S. galactic plane survey. A new H.E.S.S. source, HESS~J1908+063, which coincides with MGRO~J1908+06, is presented here. Its position, size and spectrum are measured and compared to the MILAGRO source. Possible counterparts at other wavelengths are discussed in the light of the H.E.S.S. measurements. ", "conclusions": "" }, "0710/0710.2126_arXiv.txt": { "abstract": "$\\mu$ Orionis was identified by spectroscopic studies as a quadruple star system. Seventeen high precision differential astrometry measurements of $\\mu$ Ori have been collected by the Palomar High-precision Astrometric Search for Exoplanet Systems (PHASES). These show both the motion of the long period binary orbit and short period perturbations superimposed on that caused by each of the components in the long period system being themselves binaries. The new measurements enable the orientations of the long period binary and short period subsystems to be determined. Recent theoretical work predicts the distribution of relative inclinations between inner and outer orbits of hierarchical systems to peak near 40 and 140 degrees. The degree of coplanarity of this complex system is determined, and the angle between the planes of the A-B and Aa-Ab orbits is found to be $136.7 \\pm 8.3$ degrees, near the predicted distribution peak at 140 degrees; this result is discussed in the context of the handful of systems with established mutual inclinations. The system distance and masses for each component are obtained from a combined fit of the PHASES astrometry and archival radial velocity observations. The component masses have relative precisions of $5\\%$ (component Aa), $15\\%$ (Ab), and $1.4\\%$ (each of Ba and Bb). The median size of the minor axes of the uncertainty ellipses for the new measurements is 20 micro-arcseconds ($\\microas$). Updated orbits for $\\delta$ Equulei, $\\kappa$ Pegasi, and V819 Herculis are also presented. ", "introduction": "$\\mu$ Orionis (61 Ori, HR 2124, HIP 28614, HD 40932) is a quadruple star system that has been extensively studied by radial velocity (RV) and differential astrometry. It is located just North of Betelgeuse, Orion's right shoulder (left on the sky); $\\mu$ Ori is a bright star that is visible to the unaided eye even in moderately light-polluted skies. \\cite{Frost1906} discovered it to be a short period (four and a half day) single-lined spectroscopic binary; this was component Aa, whose short-period, low mass companion Ab has never been detected directly. \\cite{Aitken1914} discovered it also had a more distant component (B) forming a sub-arcsecond visual binary. Much later, \\cite{Fekel1980} found B was itself a short-period (4.78 days) double-lined spectroscopic binary, making the system quadruple; these stars are designated Ba and Bb. Most recently, \\cite{Fekel2002} (hereafter F2002) reported the astrometric orbit of the A-B motion, double-lined RV orbits for A-B and the Ba-Bb subsystem, and a single-lined RV orbit for the Aa-Ab subsystem. F2002 estimate the spectral types as A5V (Aa, an Am star), F5V (Ba), and F5V (Bb), though they note these are classifications are less certain than usual due to the complexity of the system. For a more complete discussion of the history of $\\mu$ Ori, see F2002. Until now, astrometric observations have only been able to characterize the long period A-B motion, lacking the precision necessary to measure the astrometric perturbations to this orbit caused by the Aa-Ab and Ba-Bb subsystems. The method described by \\cite{LaneMute2004a} for ground-based differential astrometry at the $\\sim 20$ $\\microas$ level for sub-arcsecond (``speckle'') binaries has been used to study $\\mu$ Ori during the 2004-2007 observing seasons. These measurements represent an improvement in precision of over two orders of magnitude over previous work on this system. The goal of the current investigation is to report the center-of-light (photocenter) astrometric orbits of the Aa-Ab and Ba-Bb subsystems. This enables measurement of the coplanarities of the A-B, Aa-Ab, and Ba-Bb orbits. The masses and luminosity ratio of Aa and Ab are measured for the first time.Also presented are updated orbits for the PHASES targets $\\delta$ Equ, $\\kappa$ Peg, and V819 Her. Astrometric measurements were made at the Palomar Testbed Interferometer \\citep[PTI;][]{col99} as part of the Palomar High-precision Astrometric Search for Exoplanet Systems (PHASES) program \\citep{Mute06Limits}. PTI is located on Palomar Mountain near San Diego, CA. It was developed by the Jet Propulsion Laboratory, California Institute of Technology for NASA, as a testbed for interferometric techniques applicable to the Keck Interferometer and other missions such as the Space Interferometry Mission (SIM). It operates in the J ($1.2 \\mu{\\rm m}$), H ($1.6 \\mu{\\rm m}$), and K ($2.2 \\mu{\\rm m}$) bands, and combines starlight from two out of three available 40-cm apertures. The apertures form a triangle with one 110 and two 87 meter baselines. ", "conclusions": "The center-of-light astrometric motions of the Aa-Ab and Ba-Bb subsystems in $\\mu$ Ori have been constrained by PHASES observations. While four degenerate orbital solutions exist, two of these can be excluded with high reliability based on mass-luminosity arguments, and the fact that Ab is not observed in the spectra. Ba and Bb are stars of a class (mid-F dwarfs) whose properties have been well established by studying other binaries. Their association with Aa and Ab, which are members of more poorly studied classes (Am and late K dwarfs) allows a better understanding of those objects in a system which can be assumed to be coevolved. The orbital solution finds masses and luminosities for all four components, the basic properties necessary in studying their natures. Complex dynamics must occur in $\\mu$ Ori. The Ba-Bb orbital plane is nearly perpendicular to that of the A-B motion, and certainly undergoes Kozai-type inclination-eccentricity oscillations. It is possible that the mutual inclination of the A-B pair and Aa-Ab subsystem is a result of KCTF effects over the system's evolution. Finally, it is noted that the orbits in the $\\mu$ Ori system are quite non-coplanar. This is in striking contrast with the planets of the solar system, but follows the trend seen in triple star systems. With the solar system being the only one whose coplanarity has been evaluated, it is difficult to draw conclusions about the configurations of planetary systems in general. It is important that future investigations evaluate the coplanarities of extrasolar planetary systems to establish a distribution. Whether that distribution will be the same or different than that of their stellar counterparts may point to similarities or differences in star and planet formation, and provide a key constraint on modeling multiple star and planet formation." }, "0710/0710.0854_arXiv.txt": { "abstract": "{Many thermally emitting, isolated neutron stars have magnetic fields that are larger than $10^{13}$~G. A realistic cooling model that includes the presence of high magnetic fields should be reconsidered.} {We investigate the effects of an anisotropic temperature distribution and Joule heating on the cooling of magnetized neutron stars.} {The 2D heat transfer equation with anisotropic thermal conductivity tensor and including all relevant neutrino emission processes is solved for realistic models of the neutron star interior and crust.} {The presence of the magnetic field affects significantly the thermal surface distribution and the cooling history during both, the early neutrino cooling era and the late photon cooling era.} { There is a large effect of Joule heating on the thermal evolution of strongly magnetized neutron stars. Both magnetic fields and Joule heating play an important role in keeping magnetars warm for a long time. Moreover, this effect is important for intermediate field neutron stars and should be considered in radio--quiet isolated neutron stars or high magnetic field radio--pulsars.} ", "introduction": "Observation of thermal emission from neutron stars (NSs) can provide not only information on the physical properties such as the magnetic field, temperature, and chemical composition of the regions where this radiation is produced but also information on the properties of matter at higher densities deeper inside the star \\citep{Yakovlev2004,Page2006}. To derive this information, we need to calculate the structure and evolution of the star, and compare the theoretical model with the observational data. Most previous studies assumed a spherically symmetric temperature distribution. However, there is increasing evidence that this is not the case for most nearby neutron stars whose thermal emission is visible in the X-ray band of the electromagnetic spectrum \\citep{Zavlin2007,Haberl2007}. The anisotropic temperature distribution may be produced not only in the low density regions where the spectrum is formed and preliminary investigations had focused their attention, but also in intermediate density regions, such as the solid crust, where a complicated magnetic field geometry could cause a coupled magneto-thermal evolution. In some extreme cases, this anisotropy may even be present in the poorly known interior, where neutrino processes are responsible for the energy removal. The observational fact that most thermally emitting isolated NSs have magnetic fields larger than $10^{13}$ G \\citep{Haberl2007}, which is sometimes confirmed by spin down measurements, leads to the conclusion that a realistic cooling model must not avoid the inclusion of the effects produced by the presence of high magnetic fields. The transport processes that occur in the interior are affected by these strong magnetic fields and their effects are expected to have observable consequences, in particular for highly magnetized NSs or magnetars. Moreover, the large surface magnetic field strengths inferred from the observations probably indicate that the interior field could reach even larger values, as theoretically predicted by some models \\citep{TD1993}. The presence of a magnetic field affects the transport properties of all plasma components, especially the electrons. In general, the motion of free electrons perpendicular to the magnetic field is quantized in Landau levels, and the thermal and electrical conductivities exhibit quantum oscillations. In the limit of a strongly quantizing field, in which almost all electrons populate the lowest level, such as in the envelope of a NS, a quantum description is necessary to calculate the thermal and electrical conductivities. Earlier calculations by \\cite{Canuto1970} and \\cite{Itoh1975} concluded that the electron thermal conductivity is strongly suppressed in the direction perpendicular to the magnetic field and increased along the magnetic field lines, which reduces the thermal insulation of the envelope ({\\it heat blanketing}). Thus, there is an anisotropic heat transport in the NS's envelope governed by the magnetic field geometry, that produces a non-uniform surface temperature. The anisotropy in the surface temperature of a NS appears to be confirmed by the analysis of observational data from isolated NSs (see \\cite{Zavlin2007} and \\cite{Haberl2007} for reviews on the current status of theory and observations). The mismatch between the extrapolation to low energy of fits to the X-ray spectra, and the observed Rayleigh-Jeans tail in the optical band ({\\it optical excess flux}), cannot be addressed using a unique temperature. Several simultaneous fits to multiwavelength spectra of \\rxdieciocho \\citep{Pons2002,Truemper2004}, \\rbdoce \\citep{Schwope2005,Schwope2007}, and \\rxcerosiete \\citep{Perez2006} are explained by a small hot emitting area $\\simeq$ 10--20 km$^2$, and an extended cooler component. Another piece of evidence that strongly supports the nonuniform temperature distribution are pulsations in the X-ray signal of some objects of amplitudes $\\simeq$ 5--30 $\\%$, some of which have irregular light curves that point towards a non-dipolar temperature distribution. All of these facts reveal that the idealized picture of a NS with a dipolar magnetic field and uniform surface temperature is oversimplified. In a pioneering work, \\cite{Greenstein1983} obtained the temperature at the surface of a NS as a function of the magnetic field inclination angle in a simplified plane-parallel approximation. This model was applied to different magnetic field configurations and the observational consequences of a non-uniform temperature distribution were analyzed in the pulsars Vela and Geminga among others \\citep{Page1995}. \\cite{Potekhin2001} improved the former calculations including realistic thermal conductivities. Nevertheless, the temperature anisotropy as generated in the envelope may be insufficiently to be consistent with the observed thermal distribution and, in this case, should originate deeper inside the NS \\citep{Geppert2004,Azorin2006}. Crustal confined magnetic fields could be responsible for the surface thermal anisotropy. In the crust, even if a strong magnetic field is present, the electrons occupy a large number of Landau levels and the classical approximation remains valid during a long time in the thermal evolution. The magnetic field limits the movement of electrons in the direction perpendicular to the field and, since they are the main carriers of the heat transport, the thermal conductivity in this direction is highly suppressed, while remaining almost unaffected along the field lines. Temperature distributions in the crust were obtained as stationary solutions of the diffusion equation with axial symmetry \\citep{Geppert2004}. The approach assumes an isothermal core and a magnetized envelope as an inner and outer boundary condition, respectively. The results show important deviations from the crust isothermal case for crustal confined magnetic fields with strengths larger than $10^{13}$ G and temperatures below $10^{8}$ K. Similar conclusions were obtained considering not only poloidal but also toroidal components for the magnetic field \\citep{Azorin2006, Geppert2006}. This models succeeded in explaining simultaneously the observed X-ray spectrum, the optical excess, the pulsed fraction, and other spectral features for some isolated NS such as \\rxcerosiete \\citep{Perez2006} and \\rxdieciocho \\citep{Geppert2006}. Non-uniform surface temperature in NSs was studied by different authors using simplified models \\citep{ShibYak1996,Potekhin2001}. Although these models can provide useful information, a detailed investigation of heat transport in 2D must be completed to obtain more solid conclusions. However, this is not the only effect that must be revisited to study the cooling of NSs. For isolated NSs, different relevant magnetic field dissipation processes were identified \\citep{Goldreich1992}. The {\\it Ohmic} dissipation rate is determined by the finite conductivity of the constituent matter. In the crust, the electrical resistivity is mainly due to electron-phonon and electron-impurity scattering processes \\citep{Flowers1976}, resulting in more efficient Ohmic dissipation than in the fluid interior. The strong temperature dependence of the resistivity leads to rapid dissipation of the magnetic energy in the outermost low-density regions during the early evolution of a hot NS, which becomes less relevant as the star cools down. Joule heating in the crustal layers due to Ohmic decay was thought to affect only the late photon cooling era in old NS ($\\geq 10^7$ yr), and to be an efficient mechanism to maintain the surface temperature as high as $\\simeq 10^{4-5}$ K for a long time \\citep{Miralles98}. \\cite{Page2000} studied the 1-D thermal evolution of NSs combined with an evolving Stokes function that defines a purely poloidal, dipolar magnetic field. The Joule heating rate was evaluated averaging the currents over the azimuthal angle. However, for strongly magnetized NS, Joule heating can be important much earlier in the evolution. In a recent work, \\cite{Kaminker2006} placed a heat source inside the outer crust of a young, warm, magnetar of field strength $5\\times10^{14}$ G. To explain observations, they concluded that the heat source should be located at a density $\\lesssim 5\\times 10^{11}$ g~cm$^{-3}$, and the heating rate should be $\\gtrsim 10^{20}$ erg~cm$^{-3}$~s$^{-1}$ for at least $5\\times10^4$ yr. Anisotropic heat transport is neglected in these simulations, which were performed in spherical symmetry, assuming that it will not affect the results in the early evolution. Nevertheless we will show that, in 2D simulations, the effect of anisotropic heat transport is important. In addition to purely Ohmic dissipation, strongly magnetized NSs can also experience a {\\it Hall drift} with a drift velocity proportional to the magnetic field strength. Although the Hall drift conserves the magnetic energy and it is not a dissipative mechanism by itself, it can enhance the Ohmic decay by compressing the field into small scales, or by displacing currents to regions with higher resistivity, where Ohmic dissipation is more efficient. Recently, the first 2D-long term simulations of the magnetic field evolution in the crust studied the interplay of Ohmic dissipation and the Hall drift effect \\citep{PonsGeppert2007}. It was shown that, for magnetar field strength, the characteristic timescale during which Hall drift influences Ohmic dissipation is of about $10^{4}$ yr. All of these studies imply that both field decay and Joule heating play a role in the cooling of neutron stars born with field strengths $\\geq 10^{13}$ G. We will show that, during the neutrino cooling era and the early stages of the photon cooling era, the thermal evolution is coupled to the magnetic field decay, since both cooling and magnetic field diffusion proceed on a similar timescale ($\\approx 10^{6} $ yr). The energy released by magnetic field decay in the crust could be an important heat source that modifies or even controls the thermal evolution of a NS. Observational evidence of this fact is shown in \\cite{PonsLink2007}. They found a strong correlation between the inferred magnetic field and the surface temperature for a wide range of magnetic fields: from magnetars ($\\geq 10^{14}$ G), through radio-quiet isolated neutron stars ($\\simeq 10^{13}$ G) down to some ordinary pulsars ($\\leq 10^{13}$ G). The main conclusion is that, rather independently from the stellar structure and the matter composition, the correlation can be explained by the decay of currents on a timescale of $\\simeq 10^{6}$ yr. The aim of the present work is to study in a more consistent way the cooling of a realistic NS under the effects of large magnetic fields, including the effects of an anisotropic temperature distribution and Joule heating in 2D simulations. As a first step towards a fully coupled magneto-thermal evolution, a phenomenological law for the magnetic field decay is considered. This article is structured as follows. In Sect.~2 we discuss the equations governing the magnetic field structure and evolution, while Sect.~3 is devoted to the thermal evolution equations. Sect.~4 presents the microphysics inputs. Sect.~5 and 6 contain our results for weakly and strongly magnetized NSs, respectively. In Sect. 7, we focus on the effects of field decay and Joule heating on the cooling history of a NS. Finally, in Sect. 8 we present the main conclusions and perspectives of the present work. ", "conclusions": "We have presented a thorough study of the thermal evolution of neutron stars including some of the most intriguing effects of magnetic fields. Our results were based on two-dimensional cooling simulations of realistic models that account for the anisotropy in the thermal conductivity tensor. In the first part of the paper, we revisited the classical scenario with low magnetic fields and presented the input microphysics, working assumptions, and the baseline models. As an interesting byproduct, we reconsidered the growth of the crust and of the superconducting region in the NS core, and found that there are situations in which both growth rates are comparable. The main body of the work was aimed at the discussion of the two principal effects of magnetic fields: the anisotropic surface temperature distribution and the additional heating by magnetic field decay. We found that, even for purely dipolar fields, an inverted temperature distribution is plausible at intermediate ages. Thus the surface temperature distribution of neutron stars with high magnetic fields, even in the axisymmetric case, may be quite different from the model with two hot polar caps and a cooler equatorial region. The irregular light curves of some isolated neutron stars, for instance RBS1223, \\citep{Schwope2005,Kaplan2005} are an indication of such complex structures. The main result of this work is that, in NSs born as magnetars, Joule heating has an enormous effect on the thermal evolution. Moreover, this effect is important for intermediate field stars. If the magnetic field is supported by crustal currents, this effect can reach a maximum because two combined factors enhance the efficiency of the heating process: i) more heat is released into the crust, in the regions of higher resistivity close to the surface, and ii) large non radial components of the field channel the heat towards the surface, instead of being lost by neutrinos in the core. As expected, it becomes clear that magnetic fields and Joule heating are playing a key role keeping magnetars warm for a long time, but it is likely that the same effect, although quantitatively smaller, must be considered in radio--quiet isolated NSs or high magnetic field radio--pulsars. Another aspect that should be considered when we try to explain observations using theoretical cooling curves is that for many objects the age is estimated assuming that the loss of angular momentum is entirely due to dipolar radiation from a magnetic dipole (spin-down age). In the case of a decaying magnetic field, the {\\em spin down age}, seriously overestimates the {\\em true age} \\citep{GO1970}. Therefore, the cooling evolution time should be corrected, according to the prescription for magnetic field decay, to compare our model accurately with observations. A detailed comparison of the cooling curves obtained in this work with observational sources can be found in \\cite{Aguilera2008}. Our last striking remark is that the occurrence of direct URCA or, in general, fast neutrino cooling in NS may be hidden by a combination of effects due to strong magnetic fields. Our conclusion is that the most appropiate candidates to monitor as rapid coolers are NSs with fields lower than $10^{13}$ G. Otherwise, we may be misled in our interpretation of the temperature-age diagrams. The main drawback of our work is that we are not yet able to return a fully consistent simulation of the magneto-thermal coupled evolution of temperature and magnetic field. In the near future, we plan to extend this study by coupling our thermal diffusion code to the consistent evolution of the magnetic field in the crust given by the Hall induction equation. That approach will permit the accurate evaluation of the heating rates, including the non-linear effects associated with the Hall--drift in the NS crust. We believe, however, that the phenomenological parameterization employed in this paper, reproduces qualitatively the results expected in a real case. We have provided another step towards understanding the cooling of neutron stars, by pointing out a number of important features that must be more carefully considered in future work." }, "0710/0710.0406.txt": { "abstract": "We report results from a search for massive and evolved galaxies at $z \\ga 5$ in the Great Observatories Origins Deep Survey (GOODS) southern field. Combining HST ACS, VLT ISAAC and Spitzer IRAC broad--band photometric data, we develop a color selection technique to identify candidates for being evolved galaxies at high redshifts. The color selection is primarily based on locating the Balmer--break using the K- and 3.6$\\mu$m bands. Stellar population synthesis models are fitted to the SEDs of these galaxies to identify the final sample. We find 11 candidates with photometric redshifts in the range $4.9 \\leq z < 6.5$, dominated by an old stellar population, with ages 0.2$-$1.0 Gyr, and stellar masses in the range $(0.5 - 5) \\times 10^{11}$ M$_{\\odot}$. The majority of the stars in these galaxies were formed at $z > 9$ and the current star formation activity is in all cases, except two, a few percent of the inferred early star formation rate. One candidate has a spectroscopically confirmed redshift, in good agreement with our photometric redshift. The galaxies are very compact, with half--light radii in the observed $K-$band smaller than $\\sim 2$ kpc. Seven of the 11 candidates are also detected at 24$\\mu$m with the MIPS instrument on Spitzer. %%% By itself, the 24$\\mu$m emission could potentially be interpreted as PAH emission from a dusty starburst at $z \\sim 2-3$, however, it is also consistent with the presence of an obscured AGN at $z \\ga 5$. Indeed, for the $z \\ga 5$ solutions, all the MIPS detected galaxies, except two, %BBG\\#3348 and JD2, have relatively high internal extinction. While we favor the obscured AGN interpretation, based on the model SED fits to the optical/UV, we define a 'no--MIPS' sample of candidates in addition to the full sample. Results will be quoted for both samples. % We estimate the completeness of the Balmer break galaxy sample to be $\\sim$40\\% (an upper limit). The comoving number density of galaxies with a stellar mass $\\ga 10^{11}$ M$_{\\odot}$, at an average redshift $\\bar{z} = 5.2$, is $3.9 \\times 10^{-5}$ Mpc$^{-3}$ (no--MIPS sample: $1.4 \\times 10^{-5}$ Mpc$^{-3}$). The corresponding stellar mass density is $8 \\times 10^{6}$ M$_{\\odot}$ Mpc$^{-3}$ (no--MIPS sample: $6.2 \\times 10^{6}$ M$_{\\odot}$ Mpc$^{-3}$). The estimated stellar mass density at $\\bar{z} = 5.2$ is $2-3$\\% of the present day total stellar mass density and $20-25$\\% of the stellar mass density at $z \\sim 2$. If the stellar mass estimates are correct, the presence of these massive and evolved galaxies when the universe was $\\sim$1 Gyr old could suggest that conversion of baryons into stars proceeded more efficiently in the early universe than it does today. ", "introduction": "An important goal of observational cosmology is to understand how stars are assembled into galaxies and how this is related to the evolution of dark matter halos. In prevailing hierarchical models, star formation starts out in low mass systems, which build more massive galaxies through sequential merging (e.g. White \\& Rees 1978; Somerville 2004). In this picture, the most massive galaxies are found at relatively low redshifts. Recently, a significant population of galaxies with stellar mass $\\sim 10^{11}$ M$_{\\odot}$ has been found at $z \\sim 2 - 3$ (cf. Franx et al. 2003; Glazebrook et al. 2004; Fontana et al. 2004; Yan et al. 2004; Daddi et al. 2005a; Rudnick et al. 2006; van Dokkum et al. 2006). Stellar population synthesis models combined with broad-band photometric data show that many of these galaxies contain an old stellar population, with ages indicating a star formation phase within 1$-$3 Gyr after the Big Bang. Moreover, a number of submillimeter detected galaxies at $z \\sim 2 - 3$, which are known to be massive systems, based on their inferred molecular gas and dynamical mass estimates (cf. Greve et al. 2005), also appear to contain an old stellar population with mass $\\sim 10^{11}$ M$_{\\odot}$ (Borys et al. 2005). Therefore, a consensus seems to be emerging, that the most massive galaxies seen today, formed the bulk of their stars within the first $\\sim$3 Gyr of cosmic history (cf. Cimatti et al. 2004; Daddi et al. 2005b; Juneau et al. 2005). However, it is not known how these stars were assembled into their present host galaxies, whether this was done during multiple merger events, as proposed in hierarchical models, or if the stars and their host galaxy are co-eval. In view of the early formation epoch implied for many of these massive galaxies, the question whether the formation is hierarchical or monolithic becomes a matter of semantics as the merger time scale becomes comparable to the dynamical time scale. Recent ultra--deep surveys, done at wavelengths stretching from the UV to mid--infrared, have resulted in detection of galaxies and AGNs at even higher redshifts, reaching into the era of re-ionization. One example is HUDF--JD2, in the Hubble Ultra Deep Field, which Mobasher et al. (2005) identify as a candidate for a massive, evolved galaxy at $z=6.5$. The age of this galaxy is estimated to be $\\ga$600 Myrs, with a stellar mass of $\\sim 6 \\times 10^{11}$ M$_{\\odot}$, much larger than the stellar mass of the Milky Way. The implied age of this galaxy means that the bulk of the stars were formed on a short time scale just a few hundred million years after the recombination era. Other recent studies have used data from the Spitzer Space Telescope to analyze the stellar masses and ages for galaxies at $z > 5$ (cf. Yan et al. 2005, 2006; Eyles et al. 2005, 2006; Stark et al. 2006; Verma et al. 2007). The inferred stellar masses range from $1 - 10 \\times 10^{10}$ M$_{\\odot}$ and ages of several $10^8$ years. In several cases, galaxies have spectroscopically determined redshifts. Another spectroscopically confirmed galaxy is the gravitationally lensed object HCM06 at $z = 6.56$ (Hu et al. 2002), with a stellar mass of a few $10^{10}$ M$_{\\odot}$ and an age of $\\sim$300 Myr (cf. Chary, Stern \\& Eisenhardt 2005; Schaerer \\& Pell\\'{o} 2005). The presence of these massive and old galaxies at $z \\ga 5$ holds important clues for understanding how the first galaxies formed and how the galaxy population in general has evolved with cosmic time. In order to determine whether a significant population of massive and old galaxies exists at $z > 5$, and to derive the parameters characterizing this population, we need a selection method that specifically targets and selects evolved stellar systems at very high redshifts, using broad-band photometric data available from deep multiwavelength surveys. The presence of old galaxies at high redshift can not efficiently be inferred using the normal Lyman drop-out technique. The drop-out technique has proven to be efficient in detecting galaxies that are actively forming stars and contain relatively small amounts of dust, but it is ineffective for detecting galaxies without strong UV continuum. In this paper we develop a method for selecting galaxies dominated by a stellar population older than $\\sim$100 Myr and situated at $z \\ga 5$, and discuss the results and its implications. The technique is primarily based on detecting the presence of a well--developed Balmer break, redshifted to $\\sim$3$\\mu$m, that can be probed by the K$_\\mathrm{s} - 3.6\\mu$m color index. A second color index is used to further isolate the old high-z galaxies from foreground `contaminants'. The color signature of the Balmer break has previously been used to select galaxies at redshifts $z \\sim 1 - 3$ (Franx et al. 2003; Daddi et al. 2005a; Adelberger et al. 2004). By choosing a suitable filter combination, the Balmer break can be used to select galaxies at any redshift, in a manner similar to the Lyman--break technique. In this paper we will refer to the galaxies selected through this technique at $z > 5$ as Balmer--break galaxies (BBGs). The paper is structured as follows: % In Sect~\\ref{data} we present our sample and photometric data. % Sect.~\\ref{selecting} discusses the Balmer break feature, the stellar population synthesis models used and examines the confidence of the model fitting procedure using Monte Carlo simulations. In this section we also discuss degeneracies and define the final color selection criteria used in this paper. % In Sect.~\\ref{results} we present the Balmer--break candidates selected using our color criteria and model fits of synthetic stellar spectra. We derive associated physical parameters from the models and discuss the Spitzer/MIPS 24$\\mu$m detections. In this section we also discuss the individual sources and assign a confidence classification to each source based on its likelihood to have the correct redshift. % In Sect.~\\ref{testing} we apply our model fitting to galaxies with known spectroscopic redshift and assess the reliability of the estimated parameters. % In Sect.~\\ref{errors} we discuss different sources of errors and derive the completeness of our sample. % We discuss our results in Sect.~\\ref{discussion} and compare the number density of Balmer--break galaxies with the expected number density of dark matter halos. % Sect.~\\ref{summary} gives a summary of our results. % We adopt H$_0 = 72$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_m = 0.3$, and $\\Omega_{\\Lambda} = 0.7$ throughout this paper. All magnitudes are in the AB system (Oke 1974). ", "conclusions": "\\label{discussion} The existence of massive and evolved galaxies at redshifts $z \\ga 5$, when the universe was $\\la 1$ Gyr old, seems surprising at first sight. In the hierarchical scenario for galaxy formation, the majority of massive galaxies are assembled at relatively low redshifts. However, the presence of massive galaxies at high redshifts poses a fundamental problem for hierarchical models only if their number density exceeds that of correspondingly massive dark matter halos (e.g. Somerville 2004). In Sect.~\\ref{completeness} we derived the number and mass density of the Balmer--break galaxies, using our sample of 11 galaxies, as well as for a more restricted sample only containing those candidates which are not detected with MIPS at 24$\\mu$m, the 'no--MIPS sample'. %Dark matter halos By equating the co--moving number density of the Balmer--break galaxies with the expected density of dark matter halos at the same redshift, we can estimate the the maximum halo mass associated with the BBG's. Using the Sheth--Tormen modified Press--Schechter formalism (Sheth \\& Tormen 1999), with the dark matter halo concentration predicted for the revised value of the power spectrum normalization $\\sigma_{8}=0.74$ (Spergel et al. 2007), and the estimated lower limit to the number density of BBGs ($3.9 \\times 10^{-5}\\,\\eta^{-1}$ Mpc$^{-3}$), we predict a halo mass of $M_{\\mathrm{h}} = 1.0 \\times 10^{12}$ M$_{\\odot}$ (assuming the standard $\\Lambda$CDM model)\\footnote{The estimated halo mass is a non-linear function of the incompleteness coefficienct $\\eta$. For instance, with $\\eta = 0.5$, the corresponding halo mass becomes $8 \\times 10^{11}$ M$_{\\odot}$.}. Using the average stellar mass for the BBGs, we get $M_{*}/M_{\\mathrm{h}} \\approx 0.20$. Considering the no--MIPS sample, with a number density $1.4 \\times 10^{-5}\\,\\eta^{-1}$ Mpc$^{-3}$, the corresponding halo mass is $1.3 \\times 10^{12}$ M$_{\\odot}$, giving a ratio $M_{*}/M_{\\mathrm{h}} \\approx 0.08$. This estimate of the halo mass assumes that all available $\\sim 10^{12}$ M$_{\\odot}$ halos at $z \\sim 5.2$ are associated with Balmer--break galaxies. If a fraction of these halos would host lower mass stellar systems, such as Lyman--break galaxies, the $M_{*}/M_{\\mathrm{h}}$ ratio for the Balmer--break galaxies would have to increase accordingly. The ratio of the baryonic--to--total mass can be expressed in terms of a star formation efficiency parameter, $\\beta$ ($M_{*} = \\beta\\,M_{\\mathrm{baryon}}$: the fraction of baryons turned into stars over the life time of the galaxy), and the stellar mass, M$_{*}$, as, $$ M_{\\mathrm{baryon}}/M_{\\mathrm{total}} = \\beta^{-1}M_{*}/M_{\\mathrm{total}} = \\kappa $$ where $M_{\\mathrm{total}} = \\beta^{-1}\\,M_{*} + M_{\\mathrm{h}}$. We can then write $$ M_{*}/M_{\\mathrm{h}} = \\beta\\,\\kappa/(1 - \\kappa). $$ Adopting $\\kappa = 0.17$ from the WMAP3 results (Spergel et al. 2007), we get $M_{*}/M_{\\mathrm{h}} = 0.20\\,\\beta$. Klypin, Zhao \\& Somerville (2002) estimate the total (virial) and baryonic mass of the Milky Way and M31 galaxies and find $M_{*}/M_{\\mathrm{h}} = 0.06 - 0.08$, implying that in this case $\\beta = 0.3 - 0.4$. For the Balmer--break galaxies, we find $\\beta \\approx 0.4 - 1.0$, where the lower value corresponds to the no--MIPS sample. If we only consider the no--MIPS sample, the baryonic fraction is comparable to local galaxies. However, for the full sample, the results suggests that the BBGs at $z \\approx 5.2$ contain a higher fractions of baryons than galaxies at $z \\approx 0$. Another possible explanation for the high baryonic fraction is that the number density of dark matter halos at high redshift is underestimated by the Sheth--Tormen analysis, or that we have systematically overestimated the stellar masses of the BBGs by a factor $\\ga 2$. %%% Using a Chabrier or Kroupa initial mass function with a less steep low--mass end, could lower the estimated stellar masses by a factor 1.5--1.8 (see Sect.~\\ref{systematics}). %%% It would be more instructive to compare the $M_{*}/M_{\\mathrm{h}}$ ratio to that of massive elliptical galaxies at $z \\approx 0$. However, the evidence for dark matter in elliptical galaxies is still circumstancial and limited to the central regions. Using planetary nebulae and globular clusters as kinematic probes, it has been possible to push the analysis to $\\sim$5 $R_{\\mathrm{eff}}$ (e.g. Romanowsky 2003; Richtler 2004). While the number of ellipticals studied in detail is still small, the general conclusion is that most have surprisingly weak dark matter halos, i.e. large $M_{*}/M_{\\mathrm{h}}$ ratios. It remains to be determined whether the inferred $M_{*}/M_{\\mathrm{h}}$ ratio for Balmer--break galaxies is consistent with local giant elliptical galaxies. %Stellar mass density The stellar mass density of the universe from redshifts 0 to 6 has been estimated by several groups, using different samples and methods (e.g. Bell et al. 2003; Dickinson et al. 2003; Rudnick et al. 2003, 2006; Fontana et al. 2006; Yan et al. 2006). Some of these results are listed in Table~\\ref{stellarmass}, as a comparison with the results obtained for the BBGs. Most of the values listed in Table~\\ref{stellarmass} are based only on the objects observed and are lower limits. In a few cases, the mass function has been integrated to obtain the total stellar mass (e.g. Dickinson et al. 2003). In the local universe, the global stellar mass density is $(3-4) \\times 10^{8}$\\,M$_{\\odot}$ Mpc$^{-3}$, while it decreases to $\\sim 0.3 \\times 10^{8}$ M$_{\\odot}$ Mpc$^{-3}$ at $z = 2.5-3$. In Sect.~\\ref{completeness} we found that the stellar mass density of the 11 BBG candidates is $8 \\times 10^{6}\\,\\eta^{-1}$ M$_{\\odot}$ Mpc$^{-3}$. This is $\\sim 2-3$\\% of the present day total stellar mass density. Restricting the comparison to large early type galaxies in the local universe, that is, galaxies at least as massive as our BBG sample, the percentage increases to $\\sim 4-6$\\%. Comparing with the stellar mass density at $z \\sim 2$, the BBG sample already comprise $20-25$\\% of the total stellar mass found at this redshift. For the no--MIPS sample, the stellar mass density is $\\sim$5 times smaller, and in this case the comparison with stellar mass densities at lower redshifts has to be corrected accordingly. The galaxies found in this study are remarkable in that they contain a large stellar mass, have small physical sizes and that their main epoch of star formation occured at $z \\ga 10$. Galaxies with similar properties have, however, also been found by others. In a recent paper, McClure et al. (2006) searched for Lyman--break galaxies in the UKIDSS ultra deep survey, and found 9 candidates with $z > 5$. Their stellar masses are $>5 \\times 10^{10}$ M$_{\\odot}$ and their ages range from 50--500 Myr. Overall, these galaxies have properties similar to our Balmer--break galaxies. The number density for the $z > 5$ galaxies found by McClure et al. is $\\sim$4x smaller than what we find in this paper. However, the different selection process, the fact that the UKIDSS sample does not include IRAC data and the large completeness corrections needed, makes a comparison difficult. A number of massive galaxies at $z > 4$ were also found by Fontana et al. (2006) using the GOODS-MUSIC sample. Their broad--band photometric data set consists of 14 bands, including the 4 IRAC bands. The objects were identified by fitting template SED based on Bruzual \\& Charlot (2003) models to all galaxies in the sample. The best-fitting SEDs for the high redshift objects suggest that they are passively evolving galaxies, characterized by a very short time scale for star formation or by a constant star formation and a large amount of dust extinction. The stellar masses found are in excess of $10^{11}$ M$_{\\odot}$. Hence, massive and passively evolving galaxies at $z \\sim 5$ are found in several studies. A direct comparison of the results is presently not practical as different selection criteria are used, and the completeness corrections are presently poorly defined. Another surprising property of the Balmer--break galaxies is their compact sizes. As derived in Sect.~\\ref{candidateparameters}, the typical half--light radius is $\\la$2 kpc. Although this is larger than what is expected from the size vs. redshift relation derived for UV bright galaxies at similarly high redshift (Ferguson et al. 2004; Bouwens et al. 2004; Dahlen et al. in prep.), the stellar masses of the Balmer--break galaxies are at least 10 times higher. No massive compact galaxies of this type has been found in the local universe. However, compact galaxies with a large stellar mass have been found at $z \\sim 1.4$ (Trujillo et al. 2006) and at $z\\sim 2.5$ (Daddi et al. 2005b; Zirm et al. 2007; Toft et al. 2007). These galaxies are massive ($M_{*} > 10^{11}$ M$_{\\odot}$), with no sign of AGN activity and contain a passively evolving stellar population, similar to the Balmer--break galaxies. The effective radius of these galaxies, measured at rest--frame optical wavelengths, are typically $\\la 1$ kpc (Zirm et al. 2007; Toft et al. 2007), or 3--6 times smaller than local counterparts of similar stellar mass. It is hypothesized that the on--set of rapid star formation in these systems quench the star formation process, leading to very compact objects. These galaxies, as well as the Balmer--break galaxies, cannot represent fully assembled systems and must undergo subsequent evolution in their structural parameters in order to resemble local galaxies with the same stellar mass. %Implications for the reionization of the IGM The presence of a population of massive galaxies that underwent a period of intense star formation at $z \\sim 10 - 25$ is likely to have ramifications to the reionization of the intergalactic medium (IGM). Panagia et al. (2005) calculated that the star formation associated with the formation of the massive $z = 6.5$ galaxy HUDF--JD2 (Mobasher et al. 2005), could significantly contribute to the reionization of the IGM. The main uncertainties were the escape fraction of the Lyman continuum photons and the volume density of objects similar to JD2. With the new sample of post--starburst galaxies with formation redshifts in the same range as JD2, it is possible to address this question. The integrated output of Lyman continuum photons from the Balmer--break candidates depends only on the total stellar mass and the assumed IMF (Panagia et al. 2005). Because the average stellar mass for the BBG candidates is about a factor 2 smaller than for JD2, assuming the same IMF, the average number of Lyman continuum photons is likewise a factor of 2 lower. Panagia et al. (2005) concluded that if each field of 2\\ffam5 $\\times$ 2\\ffam5 contained a source like JD2, then these sources account for at least $\\sim$20\\% of the reionization of the IGM. A higher percentage is possible if the escape fraction is higher and/or the IGM is clumped. In the present case, we have a total area that is 25 times larger and a total ionizing photon output $\\sim$10 times larger than in the case of JD2. Hence, the implication is that the BBG sources can account for $\\sim$10\\% or more of the photons needed for reionization, depending on the poorly constrained parameters describing the Lyman continuum escape fraction and the clumpiness of the IGM itself. The implications for reionization are discussed in more detail in Panagia et al. (in prep.)." }, "0710/0710.1255_arXiv.txt": { "abstract": "I~have re-visited the spatial distribution of stars and high-mass brown dwarfs in the $\\sigma$~Orionis cluster ($\\sim$3\\,Ma, $\\sim$360\\,pc). The input was a catalogue of 340 cluster members and candidates at separations less than 30\\,arcmin to $\\sigma$~Ori~AB. Of them, 70\\,\\% have features of extreme youth. I~fitted the normalised cumulative number of objects counting from the cluster centre to several power-law, exponential and King radial distributions. The cluster seems to have two components: a dense core that extends from the centre to $r \\approx$ 20\\,arcmin and a rarified halo at larger separations. The radial distribution in the core follows a power-law proportional to $r^1$, which corresponds to a volume density proportional to $r^{-2}$. This is consistent with the collapse of an isothermal spherical molecular cloud. The stars more massive than 3.7\\,$M_\\odot$ concentrate, however, towards the cluster centre, where there is also an apparent deficit of very low-mass objects ($M <$ 0.16\\,$M_\\odot$). Last, I~demonstrated through Monte Carlo simulations that the cluster is azimuthally asymmetric, with a filamentary overdensity of objects that runs from the cluster centre to the Horsehead~Nebula. ", "introduction": "The $\\sigma$~Orionis region in the {Ori~OB~1~b} association is finally becoming recognised as one of the most important young open clusters, with an age of only about 3\\,Ma. In the discovery paper, Garrison (1967) used the term ``clustering'' to refer to an agglomeration of fifteen B-type stars surrounding and including the multiple star {$\\sigma$~Ori}. Afterwards, Lyng\\aa~(1981) tabulated $\\sigma$~Orionis in his catalogue of open clusters. Since the rediscovery of the cluster by Wolk (1996) and its subsequent study in depth, which has revealed the most numerous and best known substellar population (B\\'ejar et~al. 1999; Zapatero Osorio et~al. 2000, 2002; Caballero et~al. 2007), only a few authors have investigated the $\\sigma$~Orionis spatial distribution. In particular, B\\'ejar et~al. (2004) and Sherry et~al. (2004) analysed the radial distribution of $\\sigma$~Orionis cluster members and candidates in annuli of width $\\Delta r$ as a function of the separation $r$ to $\\sigma$~Ori~AB. To maximise the number of objects per annulus and minimise the Poissonian errors, $\\Delta r$ must be wide. This leads to have few annuli (no more than 12 in the $r$ = 0--30\\,arcmin interval) to fit to a suitable radial profile (exponential decay -- B\\'ejar et~al. 2004; King -- Sherry et~al.~2004). Both studies agree that the cluster may extend only up to $\\sim$25--30\\,arcmin. The low surface density of cluster members at larger separations, the sharp increase of extinction due to the nearby {Horsehead Nebula}-{Flame Nebula}-{IC~434} complex and the closeness to (or even overlapping with) other stellar populations in the Orion Belt surrounding {Alnitak} ($\\zeta$~Ori) and {Alnilam} ($\\epsilon$~Ori) prevent from suitably broaden the radial distribution analysis (Caballero 2007a). At the canonical heliocentric distance to $\\sigma$~Orionis of 360\\,pc (e.g. Brown, de~Geus \\& de~Zeeuw 1994), the cluster would have an approximate radius of~3\\,pc. In spite of the agreement on the size of $\\sigma$~Orionis, the fits and the profiles in B\\'ejar et~al. (2004) and Sherry et~al. (2004) seem to be rather incomplete and inappropiate, respectively. On the one hand, the King models were designed for tidally truncated globular clusters (King 1962, 1966; Meylan 1987), and have also been satisfactorily used for describing galaxies (e.g. Kormendy 1977; Binggeli, Sandage \\& Tarenghi 1984). These systems have had enough time to be isothermal, on the contrary to very young open clusters like $\\sigma$~Orionis, where only gravitational relaxation by initial mixing may have occurred (King 1962). On the other hand, B\\'ejar et~al. (2004) exclusively focused on the cluster substellar population. Besides, the exponential fit in B\\'ejar et~al. (2004) only accounted for the five innermost annuli, which leaded to a high uncertainty in the derived parameters. Last, in the works by Sherry et~al. (2004) and B\\'ejar et~al. (2004), the input list of cluster members and candidates came from $VRI/IZ$ optical surveys. Many sources in both analysis had no near-infrared or spectroscopic follow-up. For a correct study of the spatial distribution in $\\sigma$~Orionis, it is therefore necessary to use new fitting radial profiles and an input catalogue as comprehensive as possible. It must cover a wide mass interval. Maximum completeness and minimum contamination of the catalogue are also desired. These requirements are verified by the {\\em Mayrit} catalogue, which tabulates 339 $\\sigma$~Orionis members and candidates in a 30\\,arcmin-radius circular area centred on $\\sigma$~Ori~AB (Caballero 2007c). Of them, 241 display features of extreme youth (e.g. OB spectral types, Li~{\\sc i} in absorption, H$\\alpha$ in strong emission, spectral signatures of low gravity, near- and mid-infrared excesses due to discs). The catalogue covers three orders of magnitude in mass, from the $\\sim$18+12\\,$M_\\odot$ of the O9.5V+B0.5V binary $\\sigma$~Ori~AB to the $\\sim$0.033\\,$M_\\odot$ of the brown dwarf {B05~2.03--617} (Caballero \\& Chabrier, in~prep.). Accounting for $\\sigma$~Ori~A and B as different objects separated by $\\sim$0.25\\,arcsec, then the equatorial coordinates of 340 young stars, brown dwarfs and cluster member candidates are available. I will use this input catalogue to investigate the radial and azimuthal distribution of objects in the $\\sigma$~Orionis~region. ", "conclusions": "The $\\sim$3\\,Ma-old $\\sigma$~Orionis cluster is a perfect laboratory of star formation. I~have investigated the radial distribution of 340 cluster members and candidates in a 30\\,arcmin-radius area centred on $\\sigma$~Ori~AB, taken from Caballero (2007c). The analysis has covered a mass interval from the 18+12\\,$M_\\odot$ of $\\sigma$~Ori~AB to the $\\sim$0.03\\,$M_\\odot$ of the faintest brown dwarf detectable by DENIS. The cluster shows a clear radial density gradient, quantified by the ${\\mathcal Q}$-parameter, that accounts for the mean separation between members and the Euclidean minimum spanning tree of the cluster. I~have calculated the functional relations between normalised cumulative numbers of objects counting from the cluster centre, $f(r)$, and surface densities, $\\sigma(r)$. Cumulative distribution functions as these avoid many problems associated with binning. Among the studied radial (power-law, exponential and King) profiles, the best fit is for a composite power-law distribution of cluster members with a core and a rarified halo. The core extends up to $\\sim$20\\,arcmin from the cluster centre and is nicely modelled by a surface density $\\sigma(r) \\propto r^{-1}$, that corresponds to a volume density $\\rho(r) \\propto r^{-2}$. This volume density matches, in its turn, the radial profile in a cluster formed from the collapse of a self-gravitating, isothermal sphere. The most massive $\\sigma$~Orionis stars deviate, however, from the general trend and are much more concentrated towards the cluster centre. There is also an apparent deficit of very low-mass stars and high-mass brown dwarfs (0.16\\,$M_\\odot \\ga M \\ga$ 0.035\\,$M_\\odot$) in the innermost 4\\,arcmin and an excess in the annulus at 6--7\\,arcmin to the central Trapezium-like system. Last, there is a significant azimuthal asymmetry due to a filament-shape overdensity of objects that connects the cluster centre with a part of the Horsehead Nebula. This discovery supports the formation scenarios that predict burst of star formation in filamentary~gas." }, "0710/0710.3911_arXiv.txt": { "abstract": "% We follow the evolution of helium stars of initial mass $(2.2 - 2.5)\\,M_\\odot$, and show that they undergo off-center carbon burning, which leaves behind ${\\mathbf \\sim 0.01\\,M_\\odot}$ of unburnt carbon in the inner part of the core. When the carbon-oxygen core grows to Chandrasekhar mass, the amount of left-over carbon is sufficient to ignite thermonuclear runaway. At the moment of explosion, the star will possess an envelope of several $0.1\\,M_{\\odot}$, consisting of He, C, and possibly some H, perhaps producing a kind of peculiar SN. Based on the results of \\citet{Waldman2007} for accreting white dwarfs, we expect to get thermonuclear runaway at a broad range of $\\rho_c \\approx (1 - 6) \\times 10^9 \\mathrm{ g\\,cm^{-3}}$, depending on the amount of residual carbon. We verified the feasibility of this scenario by showing that in a close binary system with initial masses $(8.5 + 7.7)\\,M_{\\odot}$\\ and initial period of 150 day the primary produces a helium remnant of $2.3\\,M_{\\odot}$\\ that evolves further like the model we considered. ", "introduction": "Type Ia supernovae (SN~Ia) have a relatively small dispersion of luminosity (the standard deviation in peak blue luminosity is $\\sigma_B \\approx 0.4 - 0.5$ mag., \\citet{Branch1993ApJL}) and are being used as distance indicators (``standard candles''), having especial significance in the effort of determining the cosmological parameters of our universe. The long-standing explanation of the SN~Ia phenomenon has been the explosive burning of degenerate carbon in the core of a carbon-oxygen white dwarf, which becomes unstable as it grows to Chandrasekhar mass ($M_{Ch}$) either by accretion from a binary companion or by a merger of two white dwarfs, following the angular momentum loss from the system by gravitational wave radiation. However, theoretical models are still far from self-consistently producing an evolutionary path towards the progenitor and reproducing crucial features of the observational data, such as the composition of the ejecta. For a detailed review of the above see, e.g., \\citet{Leibundgut2000A&ARv, Hillebrandt&Niemeyer2000ARA&A, Filippenko2005ASSL}. As well, SN~Ia can not be regarded as perfectly homogeneous class, since their Hubble diagram exhibits scatter larger than the photometric errors, while spectroscopic and photometric peculiarities have been noted with increasing frequency in well-observed SN~Ia \\citep[e.g., ][]{Filippenko2005ASSL}. Therefore, there is an obvious need for progenitor scenarios that could explain the diversity among SN~Ia. Several explanations have been suggested, such as variations in the metallicity of the progenitor, in the carbon to oxygen ratio at its center, or in the central density at the time of ignition \\citep[e.g., ][]{Timmes2003ApJ, Ropke_etal2006A&A, Lesaffre2006MNRAS}. The variation of the latter two parameters is expected to result from the variation in the initial white dwarf mass and in the accretion history. In this work we follow the evolution of helium stars with initial mass $\\approx (2.2 - 2.5)\\,M_\\odot$ and show that they might reach thermonuclear explosion and perhaps account for some of the peculiar SNe. ", "conclusions": "" }, "0710/0710.5126_arXiv.txt": { "abstract": "Our campaign of deep monitoring observations with {\\it Chandra} of the nearby elliptical galaxy NGC 3379 has lead to the detection of nine globular cluster (GC) and 53 field low mass X-ray binaries (LMXBs) in the joint {\\it Hubble}/{\\it Chandra} field of view of this galaxy. Comparing these populations, we find a highly significant lack of GC LMXBs at the low (0.3-8~keV) X-ray luminosities (in the $\\sim 10^{36}$ to $\\sim 4\\times10^{37}$ erg s$^{-1}$ range) probed with our observations. This result conflicts with the proposition that all LMXBs are formed in GCs. This lack of low-luminosity sources in GCs is consistent with continuous LMXB formation due to stellar interactions and with the transition from persistent to transient LMXBs. The observed cut-off X-ray luminosity favors a predominance of LMXBs with main-sequence donors instead of ultra-compact binaries with white-dwarf donors; ultra-compacts could contribute significantly only if their disks are not affected by X-ray irradiation. Our results suggest that current theories of magnetic stellar wind braking may work rather better for the unevolved companions of GC LMXBs than for field LMXBs and cataclysmic variables in the Galaxy, where these companions may be somewhat evolved. ", "introduction": "Since their discovery in the Milky Way (see Giacconi 1974), the origin and evolution of Low-mass X-ray binaries (LMXBs) has been the subject of much discussion. LMXBs are found in both the stellar field and globular clusters (GCs). Their incidence per unit stellar mass is much higher in GCs than in the field, requiring a special formation mechanism, presumably dynamical (Clark 1975; Katz 1975). The high efficiency of these dynamical processes led to the suggestion that all LMXBs may form in GCs and then disperse in the field (see e.g., Grindlay 1984; Grindlay \\& Hertz 1985); however, some primordial field binaries are also expected to evolve into LMXBs, suggesting that there may be two populations of these sources, with distinct formation and evolutionary histories (see review by Verbunt \\& van den Heuvel 1995). The discovery of X-ray source populations in early-type galaxies with {\\it Chandra} has provided a wider and more diverse observational basis for the study of LMXBs, their association with GCs and the role that GCs may play in LMXB formation (see review Fabbiano 2006). However, until recently, only the most luminous extra-Galactic LMXBs with ($\\sim$0.3-8~keV) $L_X \\geq$ a few $10^{37}$erg s$^{-1}$ have been observed, and therefore the study of the GC-LMXB association has been limited to systems with X-ray luminosities in the upper range of Galactic LMXB luminosities. With our deep 337 ks {\\it Chandra} ACIS-S3 observations of the unperturbed elliptical galaxy NGC 3379 in the nearby poor group Leo (D$\\sim$11 Mpc), we can now pursue this study in a luminosity range more typical of the well-studied Galactic LMXBs. In NGC 3379, The GC-LMXB association has been previously studied by Kundu, Maccarone \\& Zepf (2007, hereafter KMZ), using the first shorter {\\it Chandra} observation of this galaxy, which has a typical source detection threshold of $\\sim 1-2 \\times 10^{37}$ erg s$^{-1}$ (KMZ). Our data set is $\\sim$10 times deeper (337 ks), allowing the detection of LMXBs at luminosities of $\\sim 10^{36}$ erg s$^{-1}$. ", "conclusions": "Our campaign of deep monitoring observations of the nearby elliptical galaxy NGC 3379 with {\\it Chandra} ACIS-S3 has led to the detection of nine GC LMXBs in the field studied optically with {\\it Hubble} WFPC2, seven of which were previously detected in a much shorter {\\it Chandra} exposure (1/10 of the exposure time; KMZ). The comparison of GC and field LMXB statistics in the joint {\\it Hubble}-{\\it Chandra} field of view demonstrates a relative lack of GC LMXB at luminosities below $\\sim 4\\times10^{37}$ erg s$^{-1}$; field and GC LMXBs instead are know to have closely matching XLFs above this luminosity (Kim E. et al 2006; KMZ). The dearth of low-luminosity GC LMXBs in NGC 3379 is consistent with a similar suggested behavior in NGC 3115 (KMZ), and with the XLFs of field and GC LMXBs in the Milky Way and M31 (Voss \\& Gilfanov 2007). These differences between low-luminosity field and GC XLFs falsify suggested theories that {\\it all} LMXBs may have been originated in GCs. The luminosity-dependent differences of field and GC XLFs cannot be explained by high luminosity GCs containing multiple LMXBs, because we find clear evidence of source variability for seven out of the nine sources invalidating this hypothesis. Persistent behavior of high luminosity GC sources, compared with transient field sources of similar high luminosity may explain the discrepancy as excess of luminous GC LMXBs. However, the detection of three candidate transient source (with peak luminosity greater than $10^{37}$ erg s$^{-1}$) in the GC LMXB population of NGC 3379 may not support this explanation. The lack of low-luminosity sources in GCs is consistent with the prediction of Bildsten and Deloye (BD4) based on the transition of sources from persistent to transients due to the thermal disk instability. However, the value of the observed XLF cut-off is not consistent with their suggestion that GC LMXBs are dominated by ultra-compact binaries, but instead favors LMXBs with H-rich MS donors. The $\\sim 4\\times10^{37}$ erg s$^{-1}$ luminosity cut-off is also consistent with current theories of magnetic stellar wind braking, suggesting that this effect may work rather better for the unevolved companions of GC LMXBs than for field LMXBs and cataclysmic variables in the Galaxy, where these companions may be somewhat evolved. While our results firmly establish a dearth of GC sources in NGC3379 at low luminosity, the accurate luminosity (and uncertainty) of the GC XLF cut-off will need the formal analysis of the NGC3379 LMXB luminosity function (Kim et al in preparation). The forthcoming analysis of the very deep {\\it Chandra} observations of NGC4278, a GC-rich galaxy (see Kim, D.-W. et al 2006), will provide in the near future stronger constraints on the potential 'universality' and value of the GC LMXB XLF cut-off luminosity." }, "0710/0710.0968_arXiv.txt": { "abstract": "{Gamma-ray binaries have been established as a new class of sources of very high energy (VHE, $>$100~GeV) photons. These binaries are composed of a massive star and a compact object. The gamma-rays are probably produced by inverse Compton scattering of the stellar light by VHE electrons accelerated in the vicinity of the compact object. The VHE emission from LS~5039 displays an orbital modulation. } {The inverse Compton spectrum depends on the angle between the incoming and outgoing photon in the rest frame of the electron. Since the angle at which an observer sees the star and electrons changes with the orbit, a phase dependence of the spectrum is expected.} {A procedure to compute anisotropic inverse Compton emission is explained and applied to the case of \\ls. The spectrum is calculated assuming the continuous injection of electrons close to the compact object: the shape of the steady-state distribution depends on the injected power-law and on the magnetic field intensity.} {Compared to the isotropic approximation, anisotropic scattering produces harder and fainter emission at inferior conjunction, crucially at a time when attenuation due to pair production of the VHE gamma-rays on star light is minimum. The computed lightcurve and spectra are very good fits to the HESS and EGRET observations, except at phases of maximum attenuation where pair cascade emission may be significant for HESS. Detailed predictions are made for a modulation in the GLAST energy range. The magnetic field intensity at periastron is 0.8$\\pm0.2$~G. } {The anisotropy in inverse Compton scattering plays a major role in \\ls. A simple model reproduces the observations, constraining the magnetic field intensity and injection spectrum. The comparison with observations, the derived magnetic field intensity, injection energy and slope suggest emission from a rotation-powered pulsar wind nebula. These results confirm gamma-ray binaries as promising sources to study the environment of pulsars on small scales.} ", "introduction": "Gamma-ray binaries have been established in the past couple of years as a new class of sources of very high energy (VHE, $>$100~GeV) photons. They are characterized by a large gamma-ray luminosity above an MeV, at the level of or exceeding their X-ray luminosity. At present, all three such systems known (\\object{\\ls}, \\object{\\psrb}\\ and \\object{\\lsi}, recently possibly joined by \\object{Cyg X-1}) comprise a massive star \\citep{2005Sci...309..746A,2005A&A...442....1A,2006Sci...312.1771A,2007arXiv0706.1505M}. The compact object in \\psrb\\ is a 48-ms, young radio pulsar. The VHE emission arises from the interaction of the relativistic wind from this pulsar, extracting rotational energy from the neutron star, with the stellar wind from its companion \\citep{1994ApJ...433L..37T}. Particles gain energy at the shock between the winds, resulting in a small-scale pulsar wind nebula \\citep{1981MNRAS.194P...1M}. The particles radiate away their energy as they are entrained in the shocked flow, forming a comet-like trail of emission behind the pulsar \\citep{2006A&A...456..801D}. The nature of the compact object and origin of the VHE emission remains controversial in \\ls\\ and \\lsi, although recent observations indicate the radio emission of \\lsi\\ behaves like the comet tail expected in the pulsar scenario \\citep{2006smqw.confE..52D}. Alternatively, the VHE emission could originate from particles accelerated in a relativistic jet, the energy source being accretion onto a black hole or neutron star \\citep{2006ApJ...643.1081D,2006A&A...451..259P}. The rationale being that there is evidence for particle acceleration in the jets of microquasars and active galactic nuclei. However, hard evidence for accretion occuring in either \\ls\\ or \\lsi\\ has been hard to come by \\citep[e.g.][]{2005A&A...430..245M} and the similarities between the three systems (and differences with the usual microquasars) do not argue in favour of the accretion/ejection scenario \\citep{2006A&A...456..801D}. Regardless of the actual powering mechanism, some particles must be accelerated to high energies to generate the VHE gamma-rays. If these particles are leptons, the only viable gamma-ray radiation mechanism is inverse Compton scattering on the stellar photons. The massive stars in gamma-ray binaries have effective temperatures of several tens of thousand K and radii of about 10~$R_\\odot$, yielding luminosities of the order of $10^{39}$~erg~s$^{-1}$. This provides a huge density of stellar photons in the UV band that VHE leptons may up-scatter, much greater than any other possible source of target photons (e.g. synchrotron or bremsstrahlung emission). The emitted VHE photons also have enough energy to produce $e^+e^-$ pairs with the UV stellar photons. Most of the VHE flux may therefore be lost to the observer if the source is behind the star and VHE photons have to travel through the stellar light. Gamma-ray attenuation has been shown to lead to a modulation of the VHE flux with minimum absorption (maximum) at inferior (superior) conjunction \\citep{2005ApJ...634L..81B,2006A&A...451....9D}. HESS observations have indeed shown a stable modulation of the VHE flux from \\ls\\ on the orbital period with a maximum around inferior conjunction. This suggests attenuation plays a role and that the source of VHE gamma-rays cannot be more than about an AU from the binary (or attenuation would be too weak to modulate the flux). However, a non-zero flux is detected at superior conjunction where a large attenuation is expected, possibly because of pair cascading. Moreover, the spectral changes that are reported do not fit with an interpretation based on pure attenuation of a constant VHE source spectrum \\citep{2006A&A...460..743A}. Inverse Compton scattering also has a well-known dependence on the angle $\\Theta$ between incoming and outgoing photon. The photon flux from the star being anisotropic, the resulting inverse Compton emission will depend on the angle at which it is observed. Hence, a phase-dependent VHE spectrum will be observed even if the distribution of particles is isotropic and remains constant throughout. This effect has previously been investigated in \\psrb\\ by \\citet{2000APh....12..335B} who calculated the radiative drag on the unshocked pulsar wind from scattering of stellar light, using results from \\citet{1989ApJ...343..277H}. The drag produces a Compton gamma-ray line with a strong dependence on viewing angle. This work purports to explain the HESS observations of LS~5039 using a combination of anisotropic inverse Compton scattering and attenuation in the simplest way possible. The aim is to constrain the underlying particle distribution and/or powering mechanism. \\S2 derives the main equations governing anisotropic Compton scattering in the context of gamma-ray binaries and discusses the principal characteristics to expect. \\S3 presents the application to the case of LS~5039. The lightcurve and spectra observed by the HESS collaboration are reproduced by a model taking into account the photon field anisotropy and the attenuation due to pair creation. \\S4 concludes on the origin of the VHE emission from this system. ", "conclusions": "The anisotropic behaviour of inverse Compton scattering has a major influence on the emission from gamma-ray binaries. In these sources, the massive star provides a large source of seed photons with energies around an electron-volt. If high energy electrons are accelerated in the vicinity of the compact object, then the angle between the star, compact object and observer changes with orbital phase. The variation in viewing angle leads to a strong modulation in both the intensity and spectral shape of the scattered radiation. Scattering stellar photons to the TeV range requires very energetic electrons with Lorentz factors $\\gamma_{\\rm e}\\approx 10^6-10^7$. The scattering therefore occurs in the Klein-Nishina regime. In this case, the anisotropy results, at inferior conjunction, in a harder and fainter spectrum than predicted using an isotropic approximation for the incoming photons. Crucially, inferior conjunction also corresponds to the phase at which the produced VHE gamma-rays are less attenuated by pair production on stellar photons. At other phases the emitted spectrum is close to the one obtained using the isotropic photon field approximation and can be severely attenuated by pair production. The result is a complex interplay that reduces the amplitude of the variations expected from a pure attenuation model and a hardening at inferior conjunction. The \\ls\\ lightcurve and spectra were modelled using a simple-minded leptonic model. The electrons are assumed to be accelerated efficiently in a small zone in the vicinity of the compact object with a standard $\\gamma_{\\rm e}^{-p}$ power-law. Radiative losses due to inverse Compton emission and synchrotron emission generate a distinctive steady-state electron distribution in this environment dominated by stellar photons. The distribution has a prominent hardening between the energy at which inverse Compton losses enter the Klein-Nishina regime ($\\gamma_{\\rm KN}\\approx 6~10^4$ in \\ls) and the energy at which synchrotron losses take over ($\\gamma_{\\rm S}\\approx 10^7$ for a 1~G field). This is for instance the distribution found in the vicinity of the pulsar wind shock but it applies equally well to any leptonic model where particles are accelerated close to the compact object. The magnetic field was allowed to vary as the inverse of the orbital separation, as expected from a pulsar wind nebula. The model has only three parameters: the intensity of the magnetic field, the normalization of the electron distribution and the slope $p$ of the injected power-law $\\gamma_{\\rm e}^{-p}$. The cutoff in the very high energy gamma-ray spectrum is very sensitive to the magnetic field intensity, via the location of $\\gamma_{\\rm S}$ in the electron distribution. Fitting the high-state spectrum seen by HESS gives a rather constrained magnetic field intensity at periastron of 0.8$\\pm0.2$~G. This value compares well with the values found using simple pulsar wind models which give 5 $(\\dot{E}_{36} \\sigma_3)^{1/2} R_{11}^{-1}$~G, where $\\dot{E}_{36}$ is the pulsar spindown power in units of 10$^{36}$~erg~s$^{1}$, $\\sigma_3$ is the ratio of magnetic to kinetic energy in the pulsar wind in units of 10$^{-3}$ and $R_{11}$ is the distance of the shock to the pulsar in units of 10$^{11}$~cm. Fitting the HESS high-state spectrum also sets the injection slope to $p=2\\pm0.3$, close to the canonical value for shock acceleration. The normalization of the electron distribution implies an injection rate of 10$^{36}$~erg~s$^{-1}$ for a radiative zone of 3~$10^{11}$~cm. These results are remarkably consistent with the expectations for a pulsar wind model. The spectrum is also found to fit extremely well the EGRET observations, adding credence to the reliability of this simple approach. The model predicts a strong variation in the GLAST band with a softening from high to low flux below a GeV (where synchrotron emission dominates the spectrum) but a hardening above a GeV (where inverse Compton emission dominates the spectrum). The HESS low-state spectrum is not explained to satisfaction. The model fits nicely the EGRET measurements but produces too many gamma-rays at 5-10~TeV. A possible solution is a more complex orbital phase-dependence of the electron distribution at selected phases. Another solution is that the low-state spectrum corresponds to phases of strong attenuation and that emission from the created pairs contribute significantly to the spectrum. Additional HESS observations near minimum flux would be welcomed. The orbital modulation of the HESS emission is easily reproduced. A well-defined peak is predicted between phases 0.7-0.9 for which evidence may already be seen in the data. The lightcurve at GLAST energies is anti-correlated with the HESS lightcurve and has a peak at periastron, where the stellar photon density is maximum, and a minimum at inferior conjunction because of the anisotropic effects in inverse Compton scattering. The GLAST spectrum below 1~GeV should be influenced by the tail of the synchrotron emission from the highest energy electrons. The peak synchrotron emission is at about $100$~MeV for maximally accelerated electrons, regardless of magnetic field. Hence, if this component is detected, it will provide evidence that electrons are indeed accelerated with extreme efficiency in this source. Similar results for the magnetic field intensity and particle energy are found when a lower inclination is used, i.e. implying a black hole compact object rather than a neutron star. In this case, the emission is thought to arise from a relativistic jet powered by accretion onto the black hole. Within the assumptions of this work on the particle distribution, it is difficult to argue that a significant part of the emission occurs far along a jet since this does not naturally reproduce neither the spectrum nor the lightcurve measured by HESS. Most of the emission should still occur close to the compact object. However, unlike in the case of a pulsar wind nebula, there is no independent theoretical expectations in support of the magnetic field intensity (certainly smaller than its equipartition value in the accretion flow) and particle energy that are derived. Therefore, the pulsar wind nebula model appears favoured independently of other possible considerations. Despite the complexity of the phenomena involved in pulsar wind nebula emission, it is found that the peculiar environment of a gamma-ray binary, most prominently the enormous luminosity of the massive companion, severely constrains the number of degrees-of-freedom in the model. A simple model suffices to reproduce most of the observations. The value of the magnetic field at the shock is found to be tightly constrained by the HESS observations to 0.8$\\pm$0.2~G and the injection spectrum slope to $p=2\\pm0.3$. These results confirm that gamma-ray binaries are promising sources to study the environment of pulsars on very small scales." }, "0710/0710.3193_arXiv.txt": { "abstract": "In this paper a neutron star with an inner core which undergoes a phase transition, which is characterized by conformal degrees of freedom on the phase boundary, is considered. Typical cases of such a phase transition are e.g. quantum Hall effect, superconductivity and superfluidity. Assuming the mechanical stability of this system the effects induced by the conformal degrees of freedom on the phase boundary will be analyzed. We will see that the inclusion of conformal degrees of freedom is not always consistent with the staticity of the phase boundary. Indeed also in the case of mechanical equilibrium there may be the tendency of one phase to swallow the other. Such a shift of the phase boundary would not imply any compression or decompression of the core. By solving the Israel junction conditions for the conformal matter, we have found the range of physical parameters which can guarantee a stable equilibrium of the phase boundary of the neutron star. The relevant parameters turn out to be not only the density difference but also the difference of the slope of the density profiles of the two phases. The values of the parameters which guarantee the stability turn out to be in a phenomenologically reasonable range. For the parameter values where the the phase boundary tends to move, a possible astrophysical consequence related to sudden small changes of the moment of inertia of the star is briefly discussed. ", "introduction": "One of the most intriguing features of neutron stars is the possibility of having an inner core which undergoes a (colour) superconductivity and/or superfluidity phase transition due to the extremely high pressure and density. This fact was first pointed out by Migdal \\cite{Mi59} (there is a huge amount of literature on this subject with a little hope of providing one with a complete list of references: for two detailed reviews, see\\ \\cite% {Pet92} \\cite{Ra99} and references therein; for general relativistic formalisms suitable for dealing with neutron stars within these scenarios see \\cite{CL98} \\cite{AC07} and references therein). Such phase transitions inside neutron stars could lead to interesting observational effects relevant in cosmology \\cite{Sigl:2006ur} and through the quasi normal modes of the stars as well as the related gravitational radiation (see, for instance, \\cite{ST02} \\cite{MP03}; detailed reviews on the subjects are \\cite% {KS99} \\cite{Ste03} \\cite{ST03}). Recently, there has been also pointed out the intriguing possibility to have a quantum hall phase of gluons and quarks in the inner core of a neutron star (see, in particular, \\cite{IM05} \\cite{Iw05}% ). The discovery of \\textit{magnetars} \\cite{DT92}, highly magnetized neutron stars whose magnetic fields can reach $10^{18}G$ in the core, makes the arising of Quantum Hall features in the inner core of a neutron star not unlikely. All the here described types of phase transitions have in common to be characterized by the presence of conformal degrees of freedom on the phase boundary predicted by QFT\\footnote{% The fact that the boundary degrees of freedom are conformal can be seen \"heuristically\" in the case of a superconducting phase. Due to the Meissner effect, the current must circulate on the phase boundary of the superconductive phase. Phenomenologically, there is no dissipation, this means that there is no physical scale over which the boundary excitations can dissipate energy. The lack of a characteristic scale for the boundary theory is related to conformal symmetry. In the case of Quantum Hall Effect, this can be easily seen from the fact that it is well described by a Chern-Simons action in the bulk. Thus, the theory induced on the boundary is the Wess-Zumino-Witten action which also has conformal symmetry.} (see, for instance, \\cite{Wi90} \\cite{We96}). The main goal of this paper is to study under which conditions the phase boundary inside a neutron star, which undergoes such a phase transition with conformal boundary degrees of freedom, can be static in general relativity, assuming mechanical equilibrium of the system\\footnote{ The analysis of the mechanical stability of neutron stars has been performed in a huge number of papers. To provide one with a complete list of references is a completely hopeless task; representative papers are, for instance, \\cite{EKO91} \\cite% {ABR00} \\cite{CL98} and references therein)}. Indeed also if the system is mechanically stable i.e. there is no pressure inducing the collapse or expansion of the core, there can be another sort of instability: there may be the tendency of one phase to prevail over the other near the boundary. This would imply that the boundary gets shifted. This situation is analogous to a bubble chamber where a small perturbation induces a phase transition. To study this phenomenon one must take into account the contribution of the nontrivial traceless boundary stress tensor to the Einstein field equations. This problem is equivalent to solving Israel's junction conditions. It is necessary to assume a dynamical phase boundary and to check under which conditions such a configuration is a solution of the Einstein field equations (by solving the junction conditions). It turns out that the equation of motion for the dynamical phase boundary is equivalent to the dynamics of a classical point particle in an effective potential. There exist a static stable equilibrium configuration if the effective potential has a local minimum for a suitable negative value of the potential. It is shown that the existence of a local minimum is regulated both by the difference in the mass density between the two phases and by difference of the slope of the two density profiles. A third parameter which determines the form of the potential is the energy of the shell. The non trivial effects of conformal boundary degrees of freedom have been pointed out, in the context of black hole physics, in \\cite% {Cao7} in which they are related to the arising (after the event horizon is formed) of Quantum Hall features \\cite{Lau99} \\cite{Wi90}\\ related to the strong attractive nature of the gravitational field acting on Fermions inside a collapsed neutron star. The structure of the paper is the following: in Section 2, the assumptions of the present paper are explained. In Section 3 the junction conditions derived from the Einstein equations for a neutron star with a phase boundary are solved. In section 4 the range of the physical parameters, characterizing our configuration, for which the phase boundary is in stable equilibrium is determined. In section 5 an astrophysical implication is discussed. Section 6 the conclusions and perspectives are drawn. ", "conclusions": "We have studied a model of a neutron star with an inner core which undergoes a phase transition. It has been shown that, even assuming mechanical equilibrium, it is not always consistent to assume a static phase boundary once conformal boundary degrees of freedom are taken into account. In order to study the staticity of such a phase boundary one must take into account the general relativistic effects of these boundary degrees of freedom by including a nontrivial stress tensor on the junction. To the best of the authors knowledge, such a consistency analysis with conformal boundary degrees of freedom has not been performed previously. The astrophysical consequences of the non-static regime, related to sudden changes of the moment of inertia of the star, are worth further investigating." }, "0710/0710.1196_arXiv.txt": { "abstract": "HI features near young star clusters in M81 are identified as the photodissociated surfaces of Giant Molecular Clouds (GMCs) from which the young stars have recently formed. The HI column densities of these features show a weak trend, from undetectable values inside $R = 3.7$\\ kpc and increasing rapidly to values around $3 \\times 10^{21}\\ \\mbox{cm}^{-2}$ near $R \\approx 7.5$\\ kpc. This trend is similar to that of the radially-averaged HI distribution in this galaxy, and implies a constant area covering factor of $\\approx 0.21$ for GMCs throughout M81. The incident UV fluxes \\Gnaught\\ of our sample of candidate PDRs decrease radially. A simple equilibrium model of the photodissociation-reformation process connects the observed values of the incident UV flux, the HI column density, and the relative dust content, permitting an independent estimate to be made of the total gas density in the GMC. Within the GMC this gas will be predominantly molecular hydrogen. Volume densities of $1 < n < 200\\ \\mbox{cm}^{-3}$ are derived, with a geometric mean of $17\\ \\mbox{cm}^{-3}$. These values are similar to the densities of GMCs in the Galaxy, but somewhat lower than those found earlier for M101 with similar methods. Low values of molecular density in the GMCs of M81 will result in low levels of collisional excitation of the CO(1-0) transition, and are consistent with the very low surface brightness of CO(1-0) emission observed in the disk of M81. ", "introduction": "\\label{sec:intro} Giant molecular clouds (GMCs) in the interstellar medium (ISM) are generally accepted as the birthplaces of new stars. The most massive of these new stars produce copious amounts of far ultraviolet (FUV) radiation which will, in turn, photodissociate the molecular gas of the parent GMCs, producing ``blankets'' of HI (and other atomic species) on their surfaces. \\citet{all1986} were the first to present evidence that major features in the HI distribution of the nearby spiral galaxy M83 (NGC 5236), namely, the inner HI spiral arms observed with the Very Large Array (VLA), were the result of photodissociation of \\Htwo on galactic scales. \\citet{all1997} confirmed that HI features existed in M81 (NGC 3031) on scales of $\\approx 150$ pc which were qualitatively consistent with the expected morphology of large, low density photodissociation regions (PDRs), and explicitly related those HI features to nearby bright sources of Far-UV (FUV) radiation found on images of the galaxy from the Ultraviolet Imaging Telescope (UIT). \\citet{smi2000} used a simple, but quantitative, model for the equilibrium physics of photodissociation regions and applied it to VLA-HI and UIT-FUV observations of M101 (NGC 5457). Their work showed that this approach provides an entirely new method for determining the volume density of molecular gas in star-forming GMCs of galaxies, a method which is no longer dependent on the (often poorly known) excitation conditions for line emission by specific molecular tracers. In this paper we return to another spiral galaxy, M81, and carry out a quantitative analysis of the GMCs in this galaxy using the methods described by \\citet{smi2000}. This new analysis has been made possible by new data on M81 which was not available to \\citet{all1997}; new, higher-resolution, high-sensitivity VLA-HI observations, and sensitive new FUV imagery from the GALEX satellite. We have also sought an independent verification that the very basis of our approach is valid, namely, that the HI in the immediate vicinity of FUV concentrations is indeed produced in PDRs. To this end we have examined the Spitzer/IRAC data on M81 for evidence of mid-IR emission by polycyclic aromatic hydrocarbons (PAHs) which are also thought to be tracers of PDRs. M81 has been probed extensively for the molecular tracer CO. No global CO map of M81 has been published to this date, and the detected CO emission is found to be very weak and spotty (see e.g. \\citet{kna2006} and references therein). Since PDRs also produce CO emission \\citep[e.g.][]{all2004}, comparing the CO results to the results in this paper is therefore of considerable interest. The outline of this paper is as follows: Section \\ref{sec:data} contains a description of the data we used, followed by a brief theoretical description of PDRs in Section \\ref{sec:theory} and the application of the method in Section \\ref{sec:method}. The results are presented in Section \\ref{sec:results}. The results are discussed and the conclusions are briefly summarized in Section \\ref{sec:discussionconclusions}. ", "conclusions": "\\label{sec:discussionconclusions} The candidate PDRs in M81, which were selected on the basis of their FUV emission, seem to fit the photodissociation model well. Our results show no systematically different properties of the parent GMCs in different parts of M81. The total hydrogen volume density is roughly constant, even as the underlying HI, FUV and dust-to-gas ratio vary. The cloud densities we find are lower than the range of values (30 - 1000 cm$^{-3}$) found in M101 \\citep{smi2000}. The observed values of $N_{HI}$ of individual PDRs are similar to those seen in M101. The observed HI columns in both galaxies abruptly increase at the same normalized radius (0.3) and appear to decline somewhat beyond 0.7, assuming an $R_{25}$ of 30.3 kpc for M101 \\citep{vil1992}. The range of observed HI columns is also the same. The downward trend in \\Gnaught\\ also is consistent with \\citet{smi2000}, when no internal extinction correction is applied to their data. Figure \\ref{fig:Rnorm_n} shows that the densities of the GMCs in M81 do not appear to change with galactocentric radius, consistent with the M101 results. \\citet{kau1999} modeled the expected CO intensity from a PDR for a range of incident FUV fluxes and cloud densities. The range of our results would be consistent with modeled CO intensities below $5 \\times 10^{-8}~ \\mbox{ergs cm}^{-2}~ \\mbox{s}^{-1}~ \\mbox{sr}^{-1}$, or $1 \\times 10^{-8}~ \\mbox{ergs cm}^{-2}~ \\mbox{s}^{-1}~ \\mbox{sr}^{-1}$ for the vast majority of sources (6.4 K km s$^{-1}$). Figure 1 in \\citet{all2004} shows that for the range of values in Figure \\ref{fig:selection}, the modeled CO intensities are independent of $n$. The low volume densities we find are consistent with a lack of CO emission in M81 as discussed by \\citet{kna2006}, and which those authors attribute to insufficient excitation. We note that the volume number densities of colliding \\Htwo\\ molecules in the GMCs of M81 is 1/2 of the values for $n$ calculated here. The mean value we have found for $n$ in the GMCs of M81 translates into a mean number density for $n_{H_2} \\approx 10 \\mbox{cm}^{-3}$, well below the values required for collisional excitation of CO molecules. In this case, the CO emission in M81 is in general \\textit{subthermal}. Their reported upper limit for $I_{CO}$ is 1.03 K km s$^{-1}$ for the regions they investigated, near our source no. 18. The \\Gnaught\\ and $n$ that we find there are consistent to that value of $I_{CO}$ (and somewhat higher) per Figure 1 in \\citet{all2004}, when beam dilution is taken into account. The low CO emission in M81 is also explored in \\citet{cas2007}, who again point to a lack of excitation of the molecular gas. They find no molecular gas in the nucleus of M81. The absence of FUV sources to excite the gas is consistent with that finding. After accounting for observational and projection effects (see the previous Section), we note that in the nearby galaxies in general on which our method is applicable, it is most sensitive to a combination of low \\Gnaught\\ and low $n$. Summarizing, our conclusions are: \\begin{itemize} \\item We selected a number of discrete FUV sources in M81, which we consider to be potential PDRs on the surfaces of the parent GMCs. \\item The total hydrogen volume densities of GMCs close to clusters of young stars in M81 are in the range of 1 $< n <$ 200 cm$^{-3}$ with a geometric mean of 17 cm$^{-3}$. This is approximately ten times lower than GMCs in M101 studied with the same method. \\item The low GMC volume densities are consistent with a lack of CO emission in M81. \\item M81's GMCs have a filling factor of $\\approx$ 0.21 within $R_{25}$. \\item No candidate PDRs are found in M81 within $R < 0.3 R_{25}$. \\item We have provided a thorough analysis of the observational selection effects on our results and conclude that, while such effects are (necessarily) present in our results, our main conclusions as to the range and values of the total volume densities of GMCs in M81 are not affected. \\item The presence of PAH emission in the neighborhood of our candidate PDRs lends support to our view that the HI patches near FUV sources are indeed produced by photodissociation. PAH emission occurs near almost all UV sources. PAH and HI emission coincide in more than half of our sources. \\end{itemize}" }, "0710/0710.1475_arXiv.txt": { "abstract": "Gamma ray bursts have been divided into two classes, long-soft gamma ray burst and short-hard gamma ray burst according to the bimodal distribution in duration time. Due to the harder spectrum and the lack of afterglows of short-hard bursts in optical and radio observations, different progenitors for short-hard bursts and long-soft bursts have been suggested. Based on the X-ray afterglow observation and the cumulative red-shift distribution of short-hard bursts, \\citet{Nak06} found that the progenitors of short-hard bursts are consistent with old populations, such as mergers of binary neutron stars. Recently, the existence of two subclasses in long-soft bursts has been suggested after considering multiple characteristics of gamma-ray bursts, including fluences and the duration time. In this work, we extended the analysis of cumulative red-shift distribution to two possible subclasses in L-GRBs. We found that two possible subclass GRBs show different red-shift distributions, especially for red-shifts $z > 1$. Our results indicate that the accumulative red-shift distribution can be used as a tool to constrain the progenitor characteristics of possible subclasses in L-GRBs. ", "introduction": "Traditionally, gamma-ray bursts (denoted as GRBs) have been divided into two classes which are separated by the duration time of gamma-ray bursts $T_{90}=2$ sec \\citep{Kou93}. Since the short-duration gamma-ray bursts had relatively hard observed spectra than long-duration gamma-ray bursts, two classes are usually called as long-soft gamma ray burst (denoted as L-GRB) and short-hard gamma-ray burst (denoted as SHB). From the afterglow observations of L-GRBs, the association between the L-GRBs and supernovae/hypernovae has been suggested. These observations suggest that the progenitors of L-GRBs are massive giant stars which are collapsing. One of the best candidate is the Collapsar model \\citep{Woo93,Mac99} in which rapidly rotating black hole with Kerr parameter about 0.8 is formed after GRB. Recently, spinning black holes with high Kerr parameters, $a_\\star \\approx 0.8$, have been observed in soft X-ray black hole binaries \\citep{McC06,Sha06}. These observations support the idea that the rapidly rotating black holes in soft X-ray black hole binaries are the remnants of L-GRBs \\citep{Bro00,Lee02}, providing the natural sources for Collapsars. Due to the differences in the observational characteristics of SHBs compared to those of L-GRBs, such as lack of afterglow observations in optical and radio band and harder spectrum, etc., different progenitors for SHBs have been suggested. Mergers of binary neutron stars \\citep{Eic89,Nar92} are among the best candidates. After recent X-ray afterglow observations by Swift and HETE-II, the associations between SHBs and the host galaxies were reported (see Table~\\ref{tab1} and references there in). Based on the X-ray afterglow observation of SHBs and the cumulative red-shift distribution of SHBs, \\citet{Nak06} found that the progenitors of SHBs are consistent with old populations. This finding supports mergers of binary neutron stars as progenitors of SHBs. Recently, the existence of two subclasses in L-GRBs has been suggested after considering both fluences and the duration time \\citep{Cha07}. They divided L-GRBs into Clusters II and III which were separated by the fluences combined with the duration time. They classified L-GRBs with higher fluences as Cluster III and suggested that their progenitors are the massive collapsing stars, such as Collapsar \\citep{Woo93,Mac99}. As a possible origin of Cluster II GRBs, they suggested neutron-star, white-dwarf binaries. If there exist two subclasses in L-GRBs as \\citet{Cha07} suggested and only Cluster III GRBs are associated with massive collapsing giants which are the progenitors of supernovae or hypernovae, the observed red-shift distribution of two subclasses might show different behaviour. In this work, we extended the work of \\citet{Nak06} to two possible subclasses in L-GRBs. We found that Clustters II and III show different red-shift distribution at high red-shift $z>1$. Our results indicate that the accumulative red-shift distribution can be used as a tool to constrain the lifetimes of possible subclasses in GRBs. In Sec.~\\ref{sec-rate}, the numerical methods which we used in our analysis are summarized \\citep{Nak06}. The estimated lifetime of SHBs are summarized in Sec.~\\ref{sec-SHB}. We confirmed that the progenitors of SHBs are consistent with old population with lifetime $\\tau_\\ast=6.5$ Gyr, indicating that SHBs are from the mergers of binary neutron stars. In Sec.~\\ref{sec-LGRB}, the cumulative red-shift distributions of two possible subclasses in L-GRB have been summarized. Our results show that there are clear differences in the red-shift distributions of two possible subclasses in L-GRBs for the red-shift $z>1$. This suggests that the cumulative red-shift distribution can be used as a tool to distinguish the subclasses in L-GRBs. Our final conclusion follows in Sec.~\\ref{sec-con}. ", "conclusions": "\\label{sec-con} In this work, by extending the work of Nakar et al. \\citep{Nak06,Nak07}, we investigated the constraints on the progenitor lifetime of SHBs. We confirmed that the SHBs are consistent with old population with mean time $\\tau_\\ast=6.5$ Gyr. We also confirmed that this conclusion is independent of the details of the cosmology models. This results support the conclusion that the progenitors of SHBs are consistent with old population, such as merging compact star binaries (neutron star--neutron star binaries or neutron star--black hole binaries) \\citep{Nak06,Nak07,Bet07,Lee07}. We also investigated two subclasses Cluster II and III L-GRBs \\citep{Cha07} using the same method of \\citet{Nak06}. We found that the power-law distribution is more consistent with Cluster III GRBs than with Cluster II GRBs. We also found that the cumulative red-shift distribution of two clusters show clear difference in the red-shift region $z>1$. This result supports the existence of two subclasses in L-GRBs. However, in order to draw firm conclusions, many effects which are not included in our analysis, e.g., the connection between GRBs and the metallicity of the host galaxy, different characteristics of cosmological GRBs and subluminous GRBs, observability premium for different type of GRB progenitors, etc., have to be considered." }, "0710/0710.1828_arXiv.txt": { "abstract": "We present high signal-to-noise spectrophotometric observations of seven luminous \\HII\\ galaxies. The observations have been made with the use of a double-arm spectrograph which provides spectra with a wide wavelength coverage, from 3400 to 10400\\,\\AA\\ free of second order effects, of exactly the same region of a given galaxy. These observations are analysed applying a methodology designed to obtain accurate elemental abundances of oxygen, sulphur, nitrogen, neon, argon and iron in the ionized gas. Four electron temperatures and one electron density are derived from the observed forbidden line ratios using the five-level atom approximation. For our best objects errors of 1\\% in t$_e$([O{\\sc iii}]), 3\\% in t$_e$([O{\\sc ii}]) and 5\\% in t$_e$([S{\\sc iii}]) are achieved with a resulting accuracy of 7\\% in total oxygen abundances, O/H. The ionisation structure of the nebulae can be mapped by the theoretical oxygen and sulphur ionic ratios, on the one side, and the corresponding observed emission line ratios, on the other -- the $\\eta$ and $\\eta$' plots --. The combination of both is shown to provide a means to test photo-ionisation model sequences currently applied to derive elemental abundances in \\HII\\ galaxies. ", "introduction": "When studying evolution two types of ages should be distinguished: the chronological and the evolutionary ages. In the case of galaxies, estimates of the chronological age can be obtained analyzing, for example, the age distribution of their stellar population while the evolutionary age can be estimated from, for example, the metal content of their interstellar medium. \\HII\\ galaxies, the subclass of Blue Compact Dwarf galaxies (BCDs) which show spectra with strong emission lines similar to those of giant extragalactic \\HII\\ regions \\citep[GEHRs;][]{1970ApJ...162L.155S,1980ApJ...240...41F}, have the lowest metal content of any starforming galaxy suggesting that they are among the youngest or less evolved galaxies known \\citep{2007ApJ...654..226R, 1972ApJ...173...25S}. After the findings that a considerable number of the objects observed at intermediate and high redshifts seem to have properties similar to the \\HII\\ galaxies we know in the Local Universe, it has been suggested that these objects might have been very common in the past and some of them may have evolved to other kind of objects \\citep{1995ApJ...440L..49K}. In order to detect these evolutionary effects we need to compare the properties of \\HII\\ galaxies both in the Local Universe and at higher redshifts. We therefore need to know the true distribution functions of their properties among which the chemical abundances are of the greatest relevance. Spectrophotometry of bright \\HII\\ galaxies in the Local Universe allows the determination of abundances from methods that rely on the measurement of emission line intensities and atomic physics. This is referred to as the \"direct\" method. In the case of more distant or intrinsically fainter galaxies, the low signal-to-noise obtained with current telescopes precludes the application of this method and empirical ones based on the strongest emission lines are required. The fundamental basis of these empirical methods is reasonably well understood \\citep[see e.g.][]{2005MNRAS.361.1063P}. The accuracy of the results however depends on the goodness of their calibration which in turn depends on a well sampled set of precisely derived abundances by the \"direct\" method so that interpolation procedures are reliable. Enlarging the calibration range is also important since, at any rate, empirically obtained relations should never be used outside their calibration validity range. The precise derivation of elemental abundances however is not a straightforward matter. Firstly, accurate measurements of the emission lines are needed. Secondly, a certain knowledge of the ionisation structure of the region is required in order to derive ionic abundances of the different elements and in some cases photoionisation models are needed to correct for unseen ionisation states. An accurate diagnostic requires the measurement of faint auroral lines covering a wide spectral range and their accurate (better than 5\\%) ratios to Balmer recombination lines. These faint lines are usually about 1\\% of the H$\\beta$ intensity. The spectral range must include from the UV [O{\\sc ii}]\\,$\\lambda$\\,3727\\,\\AA\\ doublet, to the near IR [S{\\sc iii}] $\\lambda\\lambda$\\,9069,9532\\,\\AA\\ lines. This allows the derivation of the different line temperatures: T$_e$([O{\\sc ii}]), T$_e$([S{\\sc ii}]), T$_e$([O{\\sc iii}]), T$_e$([S{\\sc iii}]), T$_e$([N{\\sc ii}]), needed in order to study the temperature and ionisation structure of each \\HII\\ galaxy considered as a multizone ionised region. Unfortunately most of the available \\HII\\ galaxy spectra have only a restricted wavelength range (usually from about 3600 to 7000\\,\\AA), consequence of observations with single arm spectrographs, and do not have the adequate S/N to accurately measure the intensities of the weak diagnostic emission lines. Even the Sloan Digital Sky Survey (SDSS; Stoughton et al.\\ 2002) \\nocite{2002AJ....123..485S} spectra do not cover simultaneously the 3727 [O{\\sc ii}] and the 9069 [S{\\sc iii}] lines, they only represent an average inside a 3\\,arcsec fibre and reach the required S/N only for the brightest objects. It is important to realise that the combination of accurate spectrophotometry and wide spectral coverage cannot be achieved using single arm spectrographs where, in order to reach the necessary spectral resolution, the wavelength range must be split into several independent observations. In those cases, the quality of the spectrophotometry is at best doubtful mainly because the different spectral ranges are not observed simultaneously. This problem applies to both objects and calibrators. Furthermore one can never be sure of observing exactly the same region of the nebula in each spectral range. To avoid all these problems the use of double arm spectrographs is required. In this work we present simultaneous blue and red observations obtained with the double arm TWIN spectrograph at the 3.5m telescope of the Spanish-German Observatory of Calar Alto. These data are of a sufficient quality as to allow the detection and measurement of several temperature sensitive lines and add to the still scarce base of precisely derived abundances. In the next section we describe some details regarding the selection of the sample as well as the observations and data reduction. The results are presented in section 3. Sections 4 and 5 are devoted to the analysis of these results which are compared with previous data in section 6. Section 7 is devoted to the discussion of our results and finally, our conclusions are summarized in section 8. \\begin{table*} \\centering \\caption[]{Journal of observations.} \\label{jour} \\begin{tabular} {@{}c c c c c c} \\hline \\multicolumn{1}{c}{Object ID} & spSpec SDSS & hereafter ID & Date & Exposure (s) & Seeing (\\arcsec)\\\\ \\hline \\uno & spSpec-52790-1351-474 & \\unoc & 2006 June 25 & 5 $\\times$ 1800 & 0.9-1.2 \\\\ \\dos & spSpec-52721-1050-274 & \\dosc & 2006 June 23 & 4 $\\times$ 1800 & 0.8-1.1 \\\\ \\tres & spSpec-52765-1293-580 & \\tresc & 2006 June 22 & 4 $\\times$ 1800 & 0.8-1.2 \\\\ \\cuatro & spSpec-52072-0617-464 & \\cuatroc & 2006 June 24 & 6 $\\times$ 1800 & 1.0-1.4 \\\\ \\cinco & spSpec-52377-0624-361 & \\cincoc & 2006 June 23 & 5 $\\times$ 1800 & 0.8-1.1 \\\\ \\seis & spSpec-52791-1176-591 & \\seisc & 2006 June 25 & 5 $\\times$ 1800 & 0.9-1.2 \\\\ \\siete & spSpec-51818-0358-472 & \\sietec & 2006 June 22 & 5 $\\times$ 1800 & 0.8-1.1 \\\\ \\hline \\end{tabular} \\end{table*} \\begin{table*} \\centering \\caption[]{Right ascension, declination, redshift and SDSS photometric magnitudes obtained using the SDSS explore tools$^a$.} \\label{obj} \\begin{tabular} {l c c c c r r c c} \\hline \\multicolumn{1}{c}{Object ID} & \\multicolumn{1}{c}{RA} & \\multicolumn{1}{c}{Dec} & redshift & \\multicolumn{1}{c}{u} & \\multicolumn{1}{c}{g} & \\multicolumn{1}{c}{r} & \\multicolumn{1}{c}{i} & \\multicolumn{1}{c}{z} \\\\ \\hline \\unoc & 14$^h$\\,55$^m$\\,06\\fs06 & 38$^{\\circ}$\\,08\\arcmin\\,16\\farcs67 & 0.028 & 18.25 & 17.57 & 17.98 & 18.23 & 18.18 \\\\ \\dosc & 15$^h$\\,09$^m$\\,09\\fs03 & 45$^{\\circ}$\\,43\\arcmin\\,08\\farcs88 & 0.048 & 18.57 & 17.72 & 18.19 & 17.87 & 17.94 \\\\ \\tresc & 15$^h$\\,28$^m$\\,17\\fs18 & 39$^{\\circ}$\\,56\\arcmin\\,50\\farcs43 & 0.064 & 18.54 & 17.88 & 18.17 & 17.52 & 17.99 \\\\ \\cuatroc & 15$^h$\\,40$^m$\\,54\\fs31 & 56$^{\\circ}$\\,51\\arcmin\\,38\\farcs98 & 0.011 & 19.11 & 18.91 & 18.97 & 19.53 & 19.46 \\\\ \\cincoc & 16$^h$\\,16$^m$\\,23\\fs53 & 47$^{\\circ}$\\,02\\arcmin\\,02\\farcs36 & 0.002 & 16.84 & 16.45 & 16.77 & 17.35 & 17.43 \\\\ \\seisc & 16$^h$\\,57$^m$\\,12\\fs75 & 32$^{\\circ}$\\,11\\arcmin\\,41\\farcs42 & 0.038 & 17.63 & 17.03 & 17.27 & 17.15 & 17.15 \\\\ \\sietec & 17$^h$\\,29$^m$\\,06\\fs56 & 56$^{\\circ}$\\,53\\arcmin\\,19\\farcs40 & 0.016 & 18.05 & 17.26 & 17.21 & 17.38 & 17.24 \\\\ \\hline \\multicolumn{9}{l}{$^a$http://cas.sdss.org/astro/en/tools/explore/obj.asp} \\end{tabular} \\end{table*} ", "conclusions": "We have performed a detailed analysis of newly obtained spectra of seven \\HII\\ galaxies selected from the Sloan Digital Sky Survey Data Release 3. The spectra cover from 3400 to 10400 \\AA\\ in wavelength at a FWHM resolution of about 2000 in the blue and 1500 in the red spectral regions. The high signal-to-noise ratio of the obtained spectra allows the measurement of four line electron temperatures: T$_e$([O{\\sc iii}]), T$_e$([S{\\sc iii}]), T$_e$([O{\\sc ii}]) and T$_e$([S{\\sc ii}]), for all the objects of the sample with the addition of T$_e$([N{\\sc ii}]) for one of the objects. These measurements and a careful and realistic treatment of the observational errors yield total oxygen abundances with accuracies between 7 and 12\\%. The fractional error is as low as 1\\% for the ionic O$ ^{2+} $/H$ ^{+} $ ratio due to the small errors associated with the measurement of the strong nebular lines of [O{\\sc iii}] and the derived T$_e$([O{\\sc iii}]), but increases to up to 30\\% for the O$^{+}$/H$^{+}$ ratio. The accuracies are lower in the case of the abundances of sulphur (of the order of 25\\% for S$^+$ and 15\\% for S$^{2+}$) due to the presence of larger observational errors both in the measured line fluxes and the derived electron temperatures. The error for the total abundance of sulphur is also larger than in the case of oxygen (between 15\\% and 20\\%) due to the uncertainties in the ionisation correction factors. This is in contrast with the unrealistically small errors quoted for line temperatures other than T$_e$([O{\\sc iii}]) in the literature, in part due to the commonly assumed methodology of deriving them from the measured T$_e$([O{\\sc iii}]) through a theoretical relation. These relations are found from photoionization model sequences and no uncertainty is attached to them although large scatter is found when observed values are plotted; usually the line temperatures obtained in this way carry only the observational error found for the T$_e$([O{\\sc iii}]) measurement and does not include the observed scatter, thus heavily understimating the errors in the derived temperature. In fact, no clear relation is found between T$_e$([O{\\sc iii}]) and T$_e$([O{\\sc ii}]) for the existing sample of objects confirming our previous results. A comparison between measured and model derived T$_e$([O{\\sc ii}]) shows than, in general, model predictions overestimate this temperature and hence underestimate the O$^+$/H$^+$ ratio. This, though not very important for high excitation objects, could be of some concern for lower excitation ones for which total O/H abundances could be underestimated by up to 0.2 dex. It is worth noting that the objects observed with double-arm spectrographs, therefore implying simultaneous and spatially coincident observations over the whole spectral range, show less scatter in the T$_e$([O{\\sc iii}])\\,-\\,T$_e$([O{\\sc ii}] plane clustering around the N$_e$\\,=\\,100 cm$^{-3}$ photo-ionisation model sequence. On the other hand, this small scatter could partially be due to the small range of temperatures shown by these objects due to possible selection effects. This small temperature range does not allow either to investigate the metallicity effects found in the ralations between the various line temperatures in recent photo-ionisation models by Izotov et al.\\ (2006). Also, the observed objects, as well as those in Paper I, though showing Ne/O and Ar/O relative abundances typical of those found for a large \\HII\\ galaxy sample (P\\'erez-Montero et al.\\ 2007), show higher than typical N/O abundance ratios that would be even higher if the [O{\\sc ii}] temperatures would be found from photo-ionisation models. We therefore conclude that approach of deriving the O$^+$ temperature from the O$^{2+}$ one should be discouraged if an accurate abundance derivation is sought. These issues could be addressed by re-observing the objects in Table \\ref{temp} , which cover an ample range in temperatures and metal content, with double arm spectrographs. This sample should be further extended to obtain a self consistent sample of about 50 objects with high S/N and excellent spectrophotometry covering simultaneously from 3600 to 9900\\,\\AA\\ This simple and easily feasible project would provide important scientific return in the form of critical tests of photoionisation models. The O$^{+} $/O$^{2+} $ and S$^{+} $/S$ ^{2+} $ ratios for all the observed galaxies, except one, cluster around a value of the ``softness parameter\" $\\eta$ of 1 implying high values of the stellar ionising temperature. For the discrepant object, showing a much lower value of $\\eta$, the intensity of the [O{\\sc ii}]\\,$\\lambda\\lambda$\\,7319,25\\,\\AA\\ lines are affected by atmospheric absorption lines. When the observational counterpart of the ionic ratios is used, this object shows a ionisation structure similar to the rest of the observed ones. This simple exercise shows the potential of checking for consistency in both the $\\eta$ and $\\eta$' plots in order to test if a given assumed ionisation structure is adequate. In fact, these consistency checks show that the stellar ionising temperatures found for the observed \\HII\\ galaxies using the ionisation structure predicted by state of the art ionisation models result too low when compared to those implied by the corresponding observed emission line ratios. Therefore, metallicity calibrations based on abundances derived according to this conventional method are probably bound to provide metallicities which are systematically too high and should be revised." }, "0710/0710.4956_arXiv.txt": { "abstract": "We present Submillimeter Array observations of the z=3.91 gravitationally lensed broad absorption line quasar \\apm\\, which spatially resolve the 1.0~mm (200~$\\mu$m rest-frame) dust continuum emission. At 0\\fasec4 resolution, the emission is separated into two components, a stronger, extended one to the northeast ($46\\pm5$~mJy) and a weaker, compact one to the southwest ($15\\pm2$~mJy). We have carried out simulations of the gravitational lensing effect responsible for the two submm components in order to constrain the intrinsic size of the submm continuum emission. Using an elliptical lens potential, the best fit lensing model yields an intrinsic (projected) diameter of $\\sim$80~pc, which is not as compact as the optical/near-infrared (NIR) emission and agrees with previous size estimates of the gas and dust emission in \\apm. Based on our estimate, we favor a scenario in which the 200~$\\mu$m (rest-frame) emission originates from a warm dust component (T$_{\\rm d}$=150-220~K) that is mainly heated by the AGN rather than by a starburst (SB). The flux is boosted by a factor of $\\sim$90 in our model, consistent with recent estimates for \\apm. ", "introduction": "\\apm\\, (=APM08) is a strongly lensed broad absorption line (BAL) quasar \\citep{irwin98,lewis98} with a very powerful active galactic nucleus \\citep[AGN;][]{soifer87}. This combination makes it an extremely bright object despite its redshift of $z$=3.91 \\citep{downes99}. Its bolometric luminosity is thought to be $\\sim$5$\\times$10$^{13}$~L$_{\\odot}$, amplified by up to a factor of 100 by a foreground galaxy which has yet to be identified. Due to the large amplification, broad absorption lines have been detected even in the X-ray \\citep{chartas02}. Ground-based, {\\it Chandra} and HST observations have suggested that \\sapm\\, consists of at least two components separated by $\\sim$$0\\farcs4$. \\citet[=I99]{ibata99} and \\citet[=E00]{egami00} later detected a third image (their image C), and argued that it is likely a third lensed image of the background QSO instead of the lensing galaxy. \\cite{lewis02b} obtained optical spectra of the three images using HST/STIS, and showed that the three spectra are quite similar. This indicates that the image C is indeed a third lensed image of \\sapm. Due to the magnification by the lens, several molecular lines have been detected in \\sapm\\, \\citep[e.g.,][] {weiss07,guelin07,riecher06,wagg06,garcia06,wagg05,lewis02a,downes99}, suggesting a molecular gas mass of M$_{\\rm gas}$(H$_2$)=8$\\times$10$^{10}$~$m^{-1}$~M$_\\odot$ \\citep[][$m$$\\equiv$magnification factor]{riecher06}. The dust mass is estimated to be M$_{\\rm dust}$=5$\\times$10$^8$~$m^{-1}$~M$_\\odot$ \\citep[][W07]{downes99,weiss07}. However, despite the increasing amount of mm-data, no tight constraints have been set on the size of the dust and/or gas emission which may help to characterize the heating source of the dust and gas, i.e., whether it is in the form of an AGN and/or starburst (SB). Only recently have W07 presented some indirect evidence through brightness temperature arguments that the molecular line and dust emission originate from a compact region with a radius of 100-200~pc. In this {\\it Letter}, we present observations from the Submillimeter Array (SMA) with sufficient resolution (0\\fasec4) to resolve the 1.0~mm dust continuum emission in \\sapm, and lensing models that constrain its intrinsic source size. ", "conclusions": "\\label{con} We have detected the 200~$\\mu$m rest-frame continuum emission in \\sapm\\, using the SMA with an angular resolution of $\\sim$0\\fasec4. The two (main) lensed images are clearly separated with a combined flux of $\\sim$60$\\pm$12~mJy. Simulations of the gravitational lensing effect in this system yield a diameter of the intrinsic (submm) continuum emitting region of $\\sim$80~pc and magnification factor of 90. Our data and simulations seem to be only consistent with a lensing scenario including a third lensed image of \\sapm\\, close to NE if the position of the compact optical/NIR emission and the position of the extended submm emission are offset from each other before lensing by $\\sim$0\\fasec003 ($\\equiv$21 pc). Further (sub)mm observations may be beneficial to determine whether such a positional offset is indeed necessary or more complicated lens potentials have to be considered. Given our size estimate, we favor a scenario in which the 200~$\\mu$m emission originates from a warm dust component that is supposed to be mainly heated by the AGN rather than by a SB." }, "0710/0710.1097_arXiv.txt": { "abstract": "We report extensive photometry of the dwarf nova V419 Lyr throughout its 2006 July superoutburst till quiescence. The superoutburst with amplitude of $\\sim 3.5$ magnitude lasted at least 15 days and was characterized by the presence of clear superhumps with a mean period of $P_{\\rm sh}=0.089985(58)$ days ($129.58\\pm0.08$ min). According to the Stolz-Schoembs relation, this indicates that the orbital period of the binary should be around 0.086 days i.e. within the period gap. During the superoutburst the superhump period was decreasing with the rate of $\\dot{P}/P_{sh}=-24.8(2.2)\\times10^{-5}$, which is one of the highest values ever observed in SU UMa systems. At the end of the plateau phase, the superhump period stabilized at a value of 0.08983(8) days. The superhump amplitude decreased from 0.3 mag at the beginning of the superoutburst to 0.1 mag at its end. In the case of V419 Lyr we have not observed clear secondary humps, which seems to be typical for long period systems. \\noindent {\\bf Key words:} Stars: individual: V419 Lyr -- binaries: close -- novae, cataclysmic variables ", "introduction": "Dwarf novae -- a subclass of Cataclysmic Variable stars -- are quite well studied interacting binary systems composed of late-type red dwarf secondary and white dwarf primary stars (Warner 1995, Hellier 2001). Matter transferred from the red dwarf forms an accretion disc around the white dwarf. Although in the last decade significant progress has been made in explaining the behaviour of dwarf novae light curves, some physical processes ongoing in these systems are still not fully understood (see for example Smak 2000, Schreiber and Lasota 2007). In particular, the thermal-tidal instability model of Osaki (1996, 2005) describing the phenomenon of superoutbursts and superhumps may be tested by examination of SU UMa-type dwarf novae light curves. Additionally, objects near and inside the so called period gap are very important from an evolutionary point of view. Those systems give us an unprecedented opportunity to study the evolution of dwarf novae. V419 Lyr is a poorly studied cataclysmic variable discovered by Kurochkin (1990) and originally classified as a Z Cam-type dwarf nova. Later, Nogami et al. (1998) caught this object in outburst and found superhumps in its light curve. Detection of superhumps together with characteristic properties of the outburst allowed them to classify V419 Lyr as a SU UMa-type dwarf nova, but short coverage of the eruption did not allow accurate determination of the superhump period. Nevertheless, there was a strong suggestion that V419 Lyr has one of the longest orbital periods known among SU UMa variables. This object has been monitored at various photometric bands by the Variable Star Network (VSNET) (see for example Kato et al. 2004a). The observations from that program enabled a tentative determination of the supercycle period to be about $\\sim 340$ days (Katysheva and Pavlenko 2003). Moreover, Morales-Rueda and Marsh (2002) obtained a spectrum of V419 Lyr during outburst showing a relatively broad absorption feature around 430-440 nm. In this work we present an analysis of photometric data collected during the 2006 July superoutburst of V419 Lyr. The data are much richer than previous studies and provide us with an opportunity to determine parameters describing this system more precisely. ", "conclusions": "The orbital period of V419 Lyr is unknown. However it is possible to estimate its value using the relation in Stolz and Schoembs (1984) connecting the period excess $\\epsilon$ defined as $P_{\\rm sh}/P_{\\rm orb}-1$ with the orbital period of the binary. This empirical relation is as follows: \\begin{equation} \\epsilon = 0.858(11) \\cdot P_{\\rm orb} - 0.0282(2) \\end{equation} Using the definition of $\\epsilon$ and knowing $P_{\\rm sh}$ for V419 Lyr, we were able to estimate the orbital period as $P_{\\rm orb}\\approx 0.086$ days. This is slightly longer than two hours which indicates that V419 Lyr is a dwarf nova in the period gap. Many characteristics of V419 Lyr are typical of SU UMa stars. It goes into superoutburst every year or so, the eruption lasts about two weeks and has an amplitude of $\\sim 3.5$ mag. Superhumps appear shortly after the beginning of the superoutburst and have a maximum amplitude of 0.3 mag, which decreases to 0.1 mag at the end of the outburst. In addition to its long orbital period, V419 Lyr is unusual in two other properties. Its superhump period derivative has one of the largest negative values known and it shows only a weak trace of secondary humps in the final stages of the superoutburst. \\bigskip \\noindent {\\bf Acknowledgments.} ~We acknowledge generous allocation of the Warsaw Observatory 0.6-m telescope time. This work used the online service of the VSNET and AAVSO. We would like to thank Prof. J\\'ozef Smak for fruitful discussions." }, "0710/0710.1574_arXiv.txt": { "abstract": "{We present the results of a {\\it Chandra} soft X-ray observation of the spectacular ionization cone in the nearby Seyfert~2 galaxy NGC~5252. As almost invariably observed in obscured AGN, the soft X-ray emission exhibits a remarkable morphological concidence with the cone ionized gas as traced by HST O[{\\sc iii}] images. Energy-resolved images and high-resolution spectroscopy suggest that the X-ray emitting gas is photoionized by the AGN, at least on scales as large as the innermost gas and stellar ring ($\\le$3~kpc). Assuming that the whole cone is photoionized by the AGN, we reconstruct the history of the active nucles in the last $\\sim$10$^5$~years.} ", "introduction": " ", "conclusions": "" }, "0710/0710.0602_arXiv.txt": { "abstract": "We describe the discovery of HAT-P-4b, a low-density extrasolar planet transiting BD+36~2593, a $V=11.2$\\,mag slightly evolved metal-rich late F star. The planet's orbital period is $3.056536\\pm0.000057$\\,d with a mid-transit epoch of $2,454,245.8154\\pm0.0003$ (HJD). Based on high-precision photometric and spectroscopic data, and by using transit light curve modeling, spectrum analysis and evolutionary models, we derive the following planet parameters: \\mpl$=0.68\\pm0.04$\\,\\mjup, \\rpl$=1.27\\pm0.05$\\,\\rjup, \\rhopl$=0.41\\pm0.06$\\,\\gcmc\\ and $a=0.0446\\pm0.0012$\\,AU\\@. Because of its relatively large radius, together with its assumed high metallicity of that of its parent star, this planet adds to the theoretical challenges to explain inflated extrasolar planets. ", "introduction": "\\label{sec:intro} In the course of our ongoing wide field planetary transit search program HATNet \\citep{bakos04}, we have discovered a large radius and low density planet orbiting an $11$th magnitude star BD+36~2593. This planet is the fifth member of a group of low-density transiting exoplanets. The combination of its low mass and the relatively high metallicity and age of the parent star makes theoretical interpretation of its large radius difficult. In this Letter we describe the observational properties of the system and derive the physical parameters both for the host star and for the planet. We also briefly comment on the theoretical status of inflated extrasolar planets. ", "conclusions": "\\label{sec:disc} We presented the discovery data and derived the physical parameters of HAT-P-4b, an inflated planet orbiting BD+36~2593. Among the 20 transiting planets discovered so far, there are five with \\rhopl~$\\lesssim0.4$\\,\\gcmc. All others have at least $50$\\% higher densities. For ease of comparison, \\tabr{puffy} lists the relevant properties of the five inflated planets. It is remarkable how similar these planets are (except for TrES-4 that has distinctively low density). With its Safronov number of $0.036$, HAT-P-4b belongs to the Class II planets according to the recent classification of \\cite{hansen07} and (together with TrES-4) further strengthens the mysterious dichotomy of the known transiting planets in this parameter. The parent star of HAT-P-4b is among the largest radii, largest mass, lowest gravity and highest metallicity transiting planet host stars. \\notetoeditor{This is the intended place of \\tabr{puffy}} \\begin{deluxetable}{lcccccc} \\tabletypesize{\\scriptsize} \\tablecaption{ Comparison of the properties of inflated planets.\\tablenotemark{a} \\label{tab:puffy}} \\tablewidth{0pt} \\tablehead{ \\colhead{Name} & \\colhead{P} & \\colhead{a} & \\colhead{M} & \\colhead{R} & \\colhead{$\\rho$} & \\colhead{$\\log g$} \\\\ \\colhead{} & \\colhead{(d)} & \\colhead{(AU)} & \\colhead{(M$_J$)} & \\colhead{(R$_J$)} & \\colhead{(\\sc{cgs})} & \\colhead{(\\sc{cgs})} } \\startdata WASP-1b & 2.52 & 0.038 & 0.87 & 1.40 & 0.39 & 3.04 \\\\ HAT-P-4b & 3.06 & 0.045 & 0.68 & 1.27 & 0.41 & 3.02 \\\\ HD~209458b & 3.53 & 0.045 & 0.64 & 1.32 & 0.35 & 2.96 \\\\ TrES-4 & 3.55 & 0.049 & 0.84 & 1.67 & 0.22 & 2.87 \\\\ HAT-P-1b & 4.47 & 0.055 & 0.53 & 1.20 & 0.38 & 2.96 \\\\ \\enddata \\tablenotetext{a} {Data from \\cite{shporer07}, this paper, \\cite{burro07}, \\cite{mandu07} and \\cite{winn07}. From top to bottom, metallicities for the parent stars are: $0.23$ \\citep{stempels07}, $0.24$, $0.02$ $0.0$ (adopted) and $0.13$.} \\end{deluxetable} Current models of irradiated giant planets are able to match the observed radii of most of the planets without invoking any additional heating mechanism. Higher metallicity cases, such as the present one, however, may pose problems (assuming that the planet and star have similar metallicities). More metals imply two opposite effects on the radius: (i) inflating it due to higher opacities in the envelope; (ii) shrinking it due to the higher molecular weight of the interior and the possible development of a large high density core. These effects have been discussed recently by \\cite{burro07}. Since WASP-1b is similar in several aspects (i.e., irradiance, metallicity) to HAT-P-4b, we consider the coreless models of WASP-1b as shown in Fig.~7 of \\cite{burro07}. It seems that HAT-P-4b can be fitted by near solar metallicity coreless models, assuming that its age is not too much greater that $4$\\,Gyr. We also refer to the layered convective mechanism of \\cite{chabrier07} that gives an alternative explanation for planets with inflated radii. We conclude that more definite statements on the relation of the observations and planet structure theories can be made only by reaching higher accuracy in the observed star/planet parameters. Nevertheless, HAT-P-4b (together with WASP-1b) does not seem to support the existence of a simple relation between host star metallicity and planet's core mass \\citep[see][]{guillot06,burro07}." }, "0710/0710.2741_arXiv.txt": { "abstract": "We study the cosmological evolution of an induced gravity model with a self-interacting scalar field $\\sigma$ and in the presence of matter and radiation. Such model leads to Einstein Gravity plus a cosmological constant as a stable attractor among homogeneous cosmologies and is therefore a viable dark-energy (DE) model for a wide range of scalar field initial conditions and values for its positive $\\gamma$ coupling to the Ricci curvature $\\gamma \\sigma^{2}R$. ", "introduction": "Several years ago a model for a varying gravitational coupling was introduced by Brans and Dicke \\cite{BD}. The model consisted of a massless scalar field whose inverse was associated with the gravitational coupling. Such a field evolved dynamically in the presence of matter and led to cosmological predictions differing from Einstein Gravity (EG) in that one generally obtained a power-law dependence on time for the gravitational coupling. In order to reduce such a strong time dependence in a cosmological setting, while retaining the Brans-Dicke results in the weak field limit, several years ago a simple model for induced gravity \\cite{CV,TV} involving a scalar field $\\sigma$ and a quartic $\\lambda\\sigma^{4}/4$ potential was introduced. This model was globally scale invariant (that is did not include any dimensional parameter) and the spontaneous breaking of scale invariance in such a context led to both the gravitational constant and inflation, through a non-zero cosmological constant \\cite{Zee}. The cosmological consequences of introducing matter as a perturbation were studied leading to a time dependence for the scalar field and consistent results. Since the present observational status is compatible with an accelerating universe dominated by dark energy (DE) we feel that the model should be studied in more detail. \\\\ In an EG framework quadratic or quartic potentials for canonical scalar fields are consistent with DE only with an extremely fine tuning in the initial conditions leading to slow-roll evolution until the present time. Indeed, a massive or a self-interacting scalar field in EG, respectively behave as dust or radiation during the coherent oscillation regime. In contrast with this, induced gravity with a self-interacting $\\lambda\\sigma^4/4$ potential has as attractor EG plus a cosmological constant on breaking scale invariance. The simple model we consider illustrates how non-trivial and non perturbative dynamics can be obtained in the context of induced gravity DE, and more generally within scalar-tensor DE. ", "conclusions": "We have shown how a simple model of self-interacting induced gravity \\cite{CV,TV} is a viable DE model, for tiny values of the self-coupling ($\\lambda \\sim {\\cal O}(10^{-128})$). The model has a stable attractor towards EG plus $\\Lambda$ and can be very similar to the $\\Lambda$CDM model for the homogeneous mode, on taking into account the Solar system constraints quoted in \\cite{Bertotti:2003rm,Eubanks,gdotref}. At the cosmological level, in the presence of radiation and dust, it is interesting that for such a simple potential (quartic for induced gravity or interacting for the equivalent Brans-Dicke model) the model has an attractor towards EG plus $\\Lambda$, very differently from the case of a massless scalar - i.e. $\\lambda = 0$ -, for which there is no mechanism of attraction towards EG. The attractor mechanism towards GR and the onset to acceleration are both inevitably triggered by the same mechanism, i.e. scale symmetry breaking in this induced gravity model. \\begin{figure}[t!!] \\centering \\epsfig{file=ws_phi4.eps, width=8.5 cm} \\caption{Evolution of $w_{\\rm DE}$ as defined in (\\ref{rhopfake}) for different choices of $\\gamma$.\\label{ws}}. \\end{figure} In contrast with EG, the choice of a runaway potential \\cite{PR} for quintessence is not mandatory (see however \\cite{others} for a study of these potential in the context of scalar-tensor theories). If the final attractor is an accelerated universe, the only constraints on parameters and initial conditions come from observations. We have shown that late time cosmology - after recombination, for instance - is mostly insensitive with respect to the initial time derivative of $\\sigma$. For this reason, the full set of parameters and initial conditions of the model are fully specified on taking the observed values for $G, H_0, \\Omega_\\Lambda$.\\\\ We have discussed in detail such model in the context of Einstein gravity, i.e. keeping the Newton constant (approximately) fixed at its actual value. The model predicts an equation of state of the equivalent Einstein model with $w_{DE}$ slightly less than $-1$ at present and homogeneous cosmology by itself seems able to constrain $\\gamma \\lesssim {\\cal O}(10^{-3})$, although a full statistical analysis is clearly beyond the scope of this work.\\\\ It is interesting to also study also what other potentials, besides the simple potential $\\lambda \\sigma^4/4$ employed in this paper, will also be compatible with the observed evolution of the universe.\\\\ \\\\ {\\bf Acknowledgment.} We wish to thank the Referee for helpful and constructive criticism." }, "0710/0710.0058_arXiv.txt": { "abstract": "The Wilkinson Microwave Anisotropy Probe (WMAP) three year results are used to constraint non-minimal inflation models. Two different non-minimally coupled scalar field potentials are considered to calculate corresponding slow-roll parameters of non-minimal inflation. The results of numerical analysis of parameter space are compared with WMAP3 data to find appropriate new constraints on the values of the non-minimal coupling. A detailed comparison of our results with previous studies reveals the present status of the non-minimal inflation model after WMAP3.\\\\ {\\bf PACS}:\\, 04.50.+h,\\, 98.80.-k\\\\ {\\bf Key Words}: Scalar-Tensor Gravity,\\, Inflation,\\, WMAP3 Data ", "introduction": "Non-minimal coupling (NMC) of scalar field with gravity is necessary in many situations of physical and cosmological interest. There are several compelling reasons to include an explicit non-minimal coupling in the action ( see for example [1,2,3,4] and references therein ). NMC arises at the quantum level when quantum corrections to the scalar field theory are considered. It is necessary also for the renormalizability of the scalar field theory in curved space. In most theories used to describe inflationary scenarios, it turns out that a non-vanishing value of the coupling constant is unavoidable [2]. In general relativity, and in all other metric theories of gravity ( in which the scalar field is not part of the gravitational sector ), the coupling constant necessarily assumes a non-vanishing value[5]. The study of the asymptotically free theories in an external gravitational field with a Gauss-Bonnet term shows a scale dependent coupling parameter. For instance, asymptotically free grand unified theories have a non-minimal coupling depending on a renormalization group parameter that converges to the value of $\\frac{1}{6}$ or to any other initial conditions depending on the gauge group and on the matter content of the theory[6]. In view of the above results and several other motivations( see for example [7,8] and references therein), it is then natural to incorporate an explicit NMC between scalar field and Ricci scalar in the inflationary paradigm. Generally, with non-minimally coupled scalar field it is harder to achieve accelerated expansion of the universe[2,7]. Part of this difficulty is related to the more sophisticated machinery of fine tuning. Nevertheless, over the last decade several non-minimal inflation scenario have been proposed to find reliable framework for issues such as graceful exit of inflationary phase and the observational constraints on the values that non-minimal coupling can attain in order to have successful inflationary scenario are discussed [8-16]. The recent astronomical observations, specially high precision data of WMAP3 [17] have opened new doors in the field of observational cosmology. As a result, these data have significant impact on the inflation paradigm. In this regard, these high precision data will give more accurate bounds on the values of non-minimal coupling in a typical non-minimal inflation model. The purpose of this letter is to study impact of WMAP3 and non-minimal inflation. Considering some well-known inflationary potentials, we explore new observational constraints on the values of non-minimal coupling to have successful non-minimal inflation. By definition of an effective scalar field potential, we show that there is a region in parameter space that inflation is driven by the non-minimal coupling term. A detailed study of spectral index and its running shows the spontaneous exit of inflationary phase ( without any mechanism ) in a suitable region of the parameter space. We also compare our results with the results of previous studies. This comparison reveals the present status of non-minimal inflation paradigm after WMAP3. ", "conclusions": "" }, "0710/0710.3167_arXiv.txt": { "abstract": "In this brief proceedings article I summarize the review talk I gave at the IAU 246 meeting in Capri, Italy, glossing over the well-known results from the literature, but paying particular attention to new, previously unpublished material. This new material includes a careful comparison of the apparently contradictory results of two independent methods used to simulate the evolution of binary populations in dense stellar systems (the direct $N$-body method of \\cite{2007ApJ...665..707H} and the approximate Monte Carlo method of \\cite{2005MNRAS.358..572I}), that shows that the two methods may not actually yield contradictory results, and suggests future work to more directly compare the two methods. ", "introduction": "Globular clusters are observed to contain significant numbers of binary star systems---so many, in fact, that they must have born with binaries (\\cite{1992PASP..104..981H}). Their presence in clusters is important for two complementary reasons. Through super-elastic dynamical scattering interactions, they act as an energy source which may postpone core collapse, and may be the dominant factor in setting the core radii of observed Galactic globulars. Similarly, the dense stellar environment and increased dynamical interaction rate in cluster cores is responsible for the high specific frequency of stellar ``exotica'' found in clusters, including low-mass X-ray binaries (LMXBs), cataclysmic variables (CVs), blue straggler stars (BSSs), and recycled millisecond pulsars (MSPs). ", "conclusions": "\\label{sec:summary} In this proceedings article I have very briefly discussed the connection between binary stars and globular cluster dynamics, moving from basic physics to current research in the span of a few paragraphs. A thorough, easily readable, and fairly recent discussion of the material can be found in \\cite{2003gmbp.book.....H}. The primary new material presented here is a comparison in phase space of the seemingly contradictory binary population evolution simulations of \\cite{2005MNRAS.358..572I} and \\cite{2007ApJ...665..707H}, showing that they may in fact both represent the same underlying physics. In other words, new simulations must be performed to better compare the two very different methods." }, "0710/0710.3351_arXiv.txt": { "abstract": "We present the results of \\spi\\ mid-infrared spectroscopic observations of two highly-obscured massive X-ray binaries: \\igrj\\ and \\gx. Our observations reveal for the first time the extremely rich mid-infrared environments of this type of source, including multiple continuum emission components (a hot component with $T$ $>$ 700~K and a warm component with $T$ $\\sim$ 180~K) with apparent silicate absorption features, numerous \\ion{H}{1} recombination lines, many forbidden ionic lines of low ionization potentials, and pure rotational \\hh\\ lines. This indicates that both sources have hot and warm circumstellar dust, ionized stellar winds, extended low-density ionized regions, and photo-dissociated regions. It appears difficult to attribute the total optical extinction of both sources to the hot and warm dust components, which suggests that there could be an otherwise observable colder dust component responsible for the most of the optical extinction and silicate absorption features. The observed mid-infrared spectra are similar to those from Luminous Blue Variables, indicating that the highly-obscured massive X-ray binaries may represent a previously unknown evolutionary phase of X-ray binaries with early-type optical companions. Our results highlight the importance and utility of mid-infrared spectroscopy to investigate highly-obscured X-ray binaries. ", "introduction": "\\label{sec_intro} Recently, a large number of highly-obscured (e.g., \\nh\\ $\\ge$ 10$^{23}$ cm$^{-2}$) massive X-ray binaries have been discovered by the \\integ\\ hard X-ray ($\\ge$ 15 keV) satellite \\citep{wet03}. The prototypical case is \\igrj\\ which shows variable high obscuration in the X-ray, sometimes reaching \\nh\\ $\\simeq$ 2 $\\times$ 10$^{24}$ \\citep{coet03, walteret03}. Its bright optical and near-infrared (IR) counterpart is an early B-type supergiant star with numerous emission lines \\citep{fc04}. Interestingly the obscuration toward \\igrj\\ obtained in the optical and near-IR wavebands (\\av\\ $\\sim$ 18) is almost two orders of magnitude smaller than that inferred from the X-rays, which suggests that the extreme obscuration seen in the X-ray is intrinsic only to the X-ray source. However, \\av\\ $\\sim$ 18 is still greater than the interstellar obscuration, indicating the existence of substantial circumstellar material around the supergiant companion. In order to fully understand the implications of the \\integ\\ discoveries --- specifically if they imply existence of a separate class of highly-obscured X-ray binaries --- we must investigate any similarities between the new \\integ\\ sources and previously known sources. As already suggested by \\citet{ret03}, the X-ray pulsar \\gx\\ appears similar to the new \\integ\\ highly-obscured massive X-ray binaries: it has variable X-ray obscuration of \\nh\\ $\\simeq$ 10$^{23}$--10$^{24}$ cm$^{-2}$, the compact source is a neutron star, and the optical companion is an early B-type supergiant \\citep[or hypergiant;][]{ket95}, like \\igrj. Recently, \\citet[][; hereafter Paper~I]{kmr06}, using optical, near-, and mid-IR ($\\le$~20~$\\mu$m) spectral energy distributions (SEDs), have found that both sources have strong mid-IR excesses that they identify as continuum emission from hot dust. This is suggestive that they have very similar circumstellar material which may be related to their strong X-ray obscuration. In this {\\em Letter}, we present the results of \\spi\\ mid-IR spectroscopic observations of \\igrj\\ and \\gx, showing that both sources indeed have very similar rich mid-IR properties previously unknown for X-ray binaries. ", "conclusions": "\\label{sec_dis_sum} Our \\spi\\ spectroscopic observations of \\igrj\\ and \\gx\\ have revealed for the first time the rich mid-IR environment of highly-obscured X-ray binaries. This includes two dust components with prominent silicate absorption, numerous \\ion{H}{1} recombination lines, many forbidden ionic lines, and pure rotational \\hh\\ lines. Based on the observed spectra, we infer the following components for \\igrj\\ and \\gx: (1) hot ($T$ $>$ 700~K) and warm ($T$ $\\sim$ 180~K) circumstellar dust; (2) ionized stellar winds responsible for the \\ion{H}{1} lines; (3) extended low-density ionized regions for the forbidden lines; and (4) photo-dissociated regions associated with the PAH, \\hh\\ and possibly the \\silii\\ line emission. For the forbidden lines, all the detected lines have relatively low ionization potentials like in starburst galaxies where the radiation is relatively soft (compared with active galactic nuclei, for instance). This may indicate that the illumination of hard X-rays from the central compact X-ray source is not primarily responsible for the forbidden line emission. However, the inferred temperature for the radiation exciting the \\neii\\ and \\neiii\\ lines is hotter than the stellar photospheres, so there can be some contribution from the compact object. Considering that \\niii\\ and \\feii\\ were detected only in \\igrj\\ while \\feiii\\ was only in \\gx, the radiation field of \\igrj\\ may be softer than \\gx, as demonstrated by the small temperature difference between the two sources (see \\S~\\ref{sec_highres}). Perhaps the most natural explanation for the origin of the hot dust component relies on dust formation in the dense outflows from the early-type companions of \\igrj\\ and \\gx, as is the case in most sgB[e] stars (i.e., B-type supergiants with forbidden emission lines). However, the origin of the warm circumstellar dust component is very uncertain, and B[e] stars seldom show evidence for this type of warm dust. If both the hot and warm dust components have spherical shell geometres around the central star, then the associated optical extinctions are: \\av\\ $\\simeq$ 4 $\\times$ 10$^{-4}$ $Q_{\\rm abs}$ ($M_{\\rm d}$/10$^{-6}$ \\msun) ($T_{\\rm d}$/100 K)$^4$ ($L_{\\rm UV}$/10$^{39}$ ergs s$^{-1}$), where $M_{\\rm d}$ and $T_{\\rm d}$ are the mass and temperature of the dust components, $Q_{\\rm abs}$ $<$ 1 is the dust absorption coefficient, and $L_{\\rm UV}$ is the ultra-violet luminosity of the central star. This gives the optical extinctions \\av\\ $\\ll$ 1 for both the hot and warm dust components of \\igrj\\ and \\gx. (Here we use 1 $\\times$ 10$^{39}$ ergs s$^{-1}$ for $L_{\\rm UV}$ for both sources.) Therefore, under the assumption of spherical shell geometry, both the hot and warm dust components of \\igrj\\ and \\gx\\ contribute very little to the total optical extinction, suggesting that the hot and warm dust components are {\\em NOT} strongly associated with the silicate absorption features. This also applies to the dust associated with the warm extended \\hh\\ gas since the optical extinction from this component is tiny (i.e., Av $\\ll$ 1; see \\S~\\ref{sec_highres}). What's then the origin of the silicate absorption features? If it is due to the foreground ISM, we would expect to see the $\\rm CO_2$ ice feature around 15 $\\mu$m, especially for \\igrj, based on the intensity ratio between the silicate absorption feature and ice feature found in the ISM \\citep{ket05}. The absence of the ice feature in our spectra supports the interpretation that the silicate absorption features are probably not associated with the foreground ISM. One possibility may be the existence of an undisclosed colder (e.g., $\\ll$ 100~K) circumstellar dust component which is responsible for the silicate absorption features and most of the optical extinction. We need further longer wavelength observatons to confirm this possibility. Considering that the 9.7~$\\mu$m silicate absorption feature represents the oxygen-rich material, the existence of the 9.7 micron silicate absorption feature may indicate that the origin of the potential colder dust component is related to the nucleosynthesis of the progenitors of \\igrj\\ and \\gx\\ (or their companions). The optical/near-IR companion of \\igrj\\ is a sgB[e] star. Such stars are known to have hot ($T$ $\\sim$ 1000~K) circumstellar dust, and probably are evolving into Luminous Blue Variables (LBVs) or Wolf-Rayet stars. Our mid-IR spectra of \\igrj\\ and \\gx\\ are very similar to that of the LBV P~Cygni which shows many \\ion{H}{1} lines and forbidden ionic lines \\citep{let96b}. The difference is that while the mid-IR continuum of P~Cygni is due to the free-free emission in stellar winds, the main mid-IR emission of \\igrj\\ and \\gx\\ is from multiple dust continua. (The SEDs of both sources observed here are inconsistent with the free-free emission, as mentioned in Paper~I.) However, we note that some LBV stars have also been observed to have thermal mid-IR dust emission \\citep{let96a}. Based on the fact that B[e] stars seldom show mid-IR forbidden line emission, the B-type supergiant (or hypergiant) companions of \\igrj\\ and \\gx\\ may be in the evolutionary track to LBVs, implying that the extremely high obscuration seen in some massive X-ray binaries may be a phenomenon associated with the evolutionary phase. This scenario is also consistent with the fact that the highly-obscured massive X-ray binary Cygnus~X-3 has a Wolf-Rayet star companion, together with the circumstellar dust emission of $T$ $\\sim$ 250~K \\citep{ket02}." }, "0710/0710.4601_arXiv.txt": { "abstract": "We constructed some main-sequence mergers from case A binary evolution and studied their characteristics via Eggleton's stellar evolution code. Both total mass and orbital angular momentum are conservative in our binary evolutions. Assuming that the matter from the secondary homogeneously mixes with the envelope of the primary and that no mass are lost from the system during the merger process, we found that some mergers might be on the left of the zero-age main sequence as defined by normal surface composition (i.e helium content $Y=0.28$ with metallicity $Z=0.02$ for Pop I) on a colour-magnitude diagram(CMD) because of enhanced surface helium content. The study also shows that central hydrogen content of the mergers is independent of mass. Our simple models provide a possible way to explain a few blue stragglers (BSs) observed on the left of zero-age main sequence in some clusters, but the concentration toward the blue side of the main sequence with decreasing mass predicted by Sandquist et al. will not appear in our models. The products with little central hydrogen in our models are probably subgiants when they are formed, since the primaries in the progenitors also have little central hydrogen and will likely leave the main sequence during merger process. As a consequence, we fit the formula of magnitude $M_{\\rm v}$ and $B-V$ of the mergers when they return back to thermal equilibrium with maximum error 0.29 and 0.037, respectively. Employing the consequences above, we performed Monte Carlo simulations to examine our models in an old open cluster NGC 2682 and an intermediate-age cluster NGC 2660. Angular momentum loss (AML) of low mass binaries is very important in NGC 2682 and its effect was estimated in a simple way. In NGC 2682, binary mergers from our models cover the region with high luminosity and those from AML are located in the region with low luminosity, existing a certain width. The BSs from AML are much more than those from our models, indicating that AML of low mass binaries makes a major contribution to BSs in this old cluster. Our models are corresponding for several BSs in NGC 2660. At the region with the most opportunity on the CMD, however, no BSs have been observed at present. {\\bf Our results are well-matched to the observations if there is $\\sim 0.5M_\\odot$ of mass loss in the merger process, but a physical mechanism for this much mass loss is a problem.} ", "introduction": "Much evidence shows that primordial binaries make an important contribution to blue stragglers (BSs) \\cite{fer03,dpa04,map04}. At present, a few BSs, i.e. F190, $\\theta$ Car, have already been confirmed to be in binaries by observations, and their formation may be interpreted by mass transfer between the components of a binary. Whereas in intermediate-age and old open and globular clusters, the number of observed close binaries among well-studied BSs is consistent with the hypothesis of binary coalescence. For example, Mateo et al. \\shortcite{mat90} made a comparison of the number of close binaries with the total number of BSs in NGC 5466 and found that it is an acceptable claim that all non-eclipsing BSs are formed as the result of mergers of the components in close binaries, though the possibility of other mechanisms to produce BSs cannot be ruled out due to the large uncertainties in their analysis. Monte-Carlo simulations of binary stellar evolution \\cite{pol94} also show that binary coalescence may be an important channel to form BSs in some clusters (e.g. with an age greater than 40 Myr). Meanwhile, the arguments in theory show that W UMa binaries (low-mass contact binaries) must eventually merge into a single star \\cite{web76,web85,ty87,mat90}. Observationally, the lack of radial velocity variations for most BSs further indicates that binary coalescence may be more important than mass transfer for BS formation \\cite{str93,pol94}. FK Comae stars are generally considered to be direct evidence for binary coalescence \\cite{str93}. The smallest mass ratio of components among observed W UMa systems to date is about 0.06. All of the above show that it is important to study the remnants of close binaries. However the merge process is complicated and the physics during the process is still uncertain. Recently, Andronov, Pinsonneault \\& Terndrup \\shortcite{apt06} studied the mergers of close primordial binaries by employing the angular momentum loss rate inferred from the spindown of open cluster stars. Their study shows that main sequence mergers can account for the observed number of single BSs in M67 and that such mergers are responsible for at least one third of the BSs in open clusters older than 1 Gyr. The physics of mergers are limiting case treatments in the study of Andronov, Pinsonneault \\& Terndrup \\shortcite{apt06}. Based on previous studies of contact binaries and some assumptions, we construct a series of merger models in this paper, to study the structure and evolution of the models and show some comparisons with observations. Case A binary evolution has been well studied by Nelson \\& Eggleton \\shortcite{nel01}. They defined six major subtypes for the evolution (AD, AR, AS, AE, AL and AN) and two rare cases (AG and AB). Three of the subtypes (AD, AR, AS) lead the binary contact as both components are main--sequence stars and two cases (AE and AG) reach contact with one or both components having left the main sequence. As there is no description for weird objects except for two merged main--sequence stars, merger products (except for two main-sequences stars) are generally assumed to have terminated their evolution \\cite{pol94}, i.e. they have left the main sequence and cannot be recognized as BSs. Here we are interested in the cases of two main-sequence stars, i.e. cases AS, AR and AD. If $t_{\\rm dyn}$, $t_{\\rm KH}$, $t_{\\rm MS}$ represent the dynamic timescale, thermal timescale and main sequence timescale of the primary (the initial massive star,*1), respectively, the following shows a simple definition of the three evolutionary cases: AD--dynamic Roche lobe overflow (RLOF), $\\dot{M}>M/t_{\\rm dyn}$; AR--rapid evolution to contact, $\\dot{M}>M/t_{\\rm KH}, t_{\\rm contact} -t_{\\rm RLOF}(*1)<0.1t_{\\rm MS}(*1)$; AS--slow evolution to contact, $t_{\\rm contact} -t_{\\rm RLOF}(*1)>0.1t_{\\rm MS}(*1)$, where $t_{\\rm RLOF}$ and $t_{\\rm contact}$ are the ages at which RLOF begins and the binary comes into contact, respectively. In case AD, the core of the secondary spirals in quickly and stays in the center of the merger. The merger then has a chemical composition similar to that of the primary, resembling the result of smoothed particle hydrodynamic calculations \\cite{lrs96,sill97,sill01}. We therefore studied just the systems in cases AR and AS for this work. \\section {Assumptions} Using the stellar evolution code devised by Eggleton \\shortcite{egg71,egg72,egg73}, which has been updated with the latest physics over the last three decades \\cite{han94,pol95,pol98}, we re-calculate the models of cases AS and AR with primary masses between 0.89 and $2M_\\odot$ until the systems become contact binaries. The structures of the primaries and the compositions of the secondaries are stored to construct the merger remnants. Before the system comes into contact, the accreting matter is assumed to be deposited onto the surface of the secondary with zero falling velocity and distributed homogeneously all over the outer layers. The change of chemical composition on the secondary's surface caused by the accreting matter is \\begin{equation} {\\partial X_i / \\partial t }={(\\partial M /\\partial t)/[(\\partial M /\\partial t){\\rm d}t+M_{\\rm s}}] \\cdot (X_{i{\\rm a}}-X_{i{\\rm s}}), \\end {equation} where $\\partial M /\\partial t$ is the mass accretion rate, $X_{i{\\rm a}}$ and $X_{i{\\rm s}}$ are element abundances of the accreting matter and of the secondary's surface for species $i$, respectively, and $M_{\\rm s}$ is the mass of the outermost layer of the secondary. The value of $M_{\\rm s}$ will change with the moving of the non-Lagrangian mesh as well as the chosen model resolution, but it is so small ($\\sim 10^{-9}-10^{-12} M_{\\odot}$) in comparison with $(\\partial M /\\partial t){\\rm d}t$ ($\\sim 10^{-3}-10^{-5} M_{\\odot}$) during RLOF that we may ignore the effect of various $M_{\\rm s}$ on element abundances. Before and after RLOF, we get $\\partial X_{\\rm i}/\\partial t =0$ from the equation, which is reasonable in the absence of mixing \\cite{ch04}. The merger models are constructed based on the following assumptions: (i) contact binaries with two main-sequence components coalesce finally and the changes of structures of individual components during coalescence are ignored; (ii) the matter of the secondary is homogeneously mixed with that of the primary beyond the core-envelope transition point, which separates the core and the envelope of the mass donor; (iii) the system mass is conserved. Firstly, we present a brief discussion on these assumptions. Webbink \\shortcite{web76} studied the evolutionary fate of low-mass contact binaries, and found that a system cannot sustain its binary character beyond the limits set by marginal contact evolution ($\\mu =M_1/(M_1+M_2)=1.0$). He stated that a contact binary will very likely coalesce as the primary is still on the main sequence in a real system. Up to now, it is widely believed that case AD probably leads to common envelope, spiral-in, and coalescence on quite a short timescale. The final consequences of AS and AR are not very clear, but Eggleton \\shortcite{egg00} pointed out that systems undergoing AR or AS evolution may maintain a shallow contact (perhaps intermittently) as the mass ratio becomes more extreme, and finally coalesce. Recent study on W UMa (Li, Zhang \\& Han, 2005) also shows that these systems will be eventually coalescence. The merged timescale, i.e. the time from a binary contact to coalescence, is important here. If it is too long, the structures of both components will change remarkably and the system may have not completed coalescence within the cluster age. There are many {\\bf conflicting estimates} for the timescale, however, from observations and theoretical models of the merger process. {\\bf Early observational estimates range from} $10^7$--$10^8$ yr in various environments \\cite{van79,eggen89}. The following study explored the average age about $5 \\times 10^8$ yr \\cite{van94,dry02}. Bilir et al. \\shortcite{bil05} pointed out that the age difference between field contact binaries and chromospherically active binaries, 1.61 Gyr, is likely an upper limit for the contact stage by assuming an equilibrium in the Galaxy, whereas the study of W UMa by Li, Han \\& Zhang \\shortcite{lhz04} suggested a much longer timescale, about 7 Gyr. We adopt the empirically estimated values in this paper {\\bf (i.e from $5 \\times 10^7$ to $1 \\times 10^9$ yrs)} and ignore the changes of structure of individual components during merger process. For low-mass contact binaries, the common envelope is convective \\cite{web77}, and the matter in it is thus homogeneous. If a system mimics shallow contact during coalescence, it is reasonable to assume that the matter of the secondary mixes with the envelope homogeneously. Van't Veer \\shortcite{van97} found that the mass loss from the system during coalescence is at a rate of about $2 \\times 10^{-10}M_\\odot{\\rm yr}^{-1}$ by observations. If we consider that the coalescence time is $5 \\times 10^8$ yr in a binary, only $0.1M_\\odot$ is lost from the system as the binary finally becomes a single star. We then roughly assume that the mass is conservative during coalescence. However mass loss might be an important way to carry orbital angular momentum away from the binary in this process. Secondly, we discuss the choice of the core-envelope transition point which separates the core and the envelope in the primary. Many characteristics of the merger are relevant to the choice, e.g. the chemical composition in the envelope, evolutionary track on Hertzsprung-Russel diagram, and some observational characteristics. Unfortunately, one cannot find the core-envelope transition point in a main-sequence star as easily as in evolved stars because the density profile, as well as many other thermodynamic quantities (entropy, pressure, temperature etc.), is smooth and does not have a deep gradient for main-sequence stars. Chen \\& Han \\shortcite{ch05} studied the influences of core-envelope transition point on the mergers of contact binaries with two main-sequence components. They found that one may ignore the effects which result from different choices of the transition point on colours and magnitudes of the merger if it is outside the nuclear reaction region of the primary, which is commonly considered as the nearest boundary of the secondary reaches in cases AS and AR. In this paper, the core-envelope transition is determined as the point within which the core produces 99 per cent of total luminosity. This choice is generally outside the nuclear reaction regions and has little effect on the final results. Finally the merger remnant is constructed as follows: it has the total mass of the system and a chemical composition within $M_{\\rm 1c}$ similar to the core of the primary. The chemical composition in the envelope of the merger is given by \\begin{equation} X_i=(M_{i2}+M_{i1\\rm b})/(M_2+m_{\\rm b}), \\end{equation} where $M_{i2}$ and $M_{i1\\rm b}$ denote the total masses of species $i$ of the secondary and of the primary's envelope, respectively. $m_{\\rm b}$ is the envelope mass of the primary. There might be a region in which the helium abundance is less than that of the outer region. The matter in this region then has a lower mean molecular weight than that in the outer region. This results in secular instability and thermohaline mixing \\cite{kip80,ulr72}. We include it as a diffusion process in our code \\cite{ch04}. In the models of Nelson \\& Eggleton \\shortcite{nel01}, both total mass and angular momentum are conservative. It was mentioned by the authors, however, that these assumptions were only reasonable for a restricted range of intermediate masses, i.e spectra from about G0 to B1 and luminosity class III-V. Observationally, some low mass binaries with late-type components show clear signs of magnetic activity, which indicates that the systems evolve by way of a scenario implying angular momentum loss (AML) by magnetic braking \\cite{mes84}. Magnetized stellar winds probably do not carry off much mass, but they are rich in angular momentum because of magnetic linkage to the binaries. For close binaries, rotation is expected to synchronize with orbital period, so AML is at the expense of the orbital angular momentum, resulting in orbital decaying and the components approaching each other. A detached binaries, then, may become contact and finally coalesce at or before the cluster age \\cite{ste95}. There are a number of subjects including the treatment of AML \\cite{lhz04,ste06,mk06,dek06}. For simplicity, the conservative assumption is also adopted in our binary evolutions. In old clusters, however, AML of low mass binaries is very important and {\\bf we estimate its importance in another way (see section 4.2).} ", "conclusions": "In Sect.4, we notice that the timescale from contact to complete coalescence, $t_{\\rm cc}$, strongly affects the initial parameter space of primordial binaries which eventually produce single BSs in a cluster. On the other hand, there are {\\bf some conflicting estimates} for $t_{\\rm cc}$ from observations and theoretical models. {\\bf In this paper we adopt empirical values, i.e. $t_{\\rm cc}$ is short in comparison to the evolution timescale of both components in a binary, and ignore the changes during the merger process. In this section, we will first discuss the consequences of a long $t_{\\rm cc}$. Because of evolution of both components during the merger process, the primaries have lower central hydrogen content and the matter from the secondaries have larger He content. The former results in a redder colour for the mergers while the latter makes the mergers bluer. So the final positions of the mergers are possibly similar to those shown in this paper, except that the primaries have left the main sequence at final coalescence. This case will appear in the mergers with little central hydrogen. For example, a star with $2M_\\odot$ may evolve from ZAMS to exhausted of central hydrogen in $10^9$ yr, and then none of the mergers from binaries with primary' masses larger than $2M_\\odot$ will be on the main sequence if $t_{\\rm cc}= 1 \\times 10^9$ yr. For the primaries with very little hydrogen in the center at contact, the mergers may never be on the main sequence even in the cases of short $t_{\\rm cc}$. The long $t_{\\rm cc}$ also delays the appearance of the mergers and shortens their timescales on the main sequence. The latter has not been exactly expressed in our models, and therefore we just see that the mergers from a long $t_{\\rm cc}$ have larger luminosities as shown in section 4.} {\\bf In our binary evolutions, we have not included AML, which exists in low-mass binaries and may be the main course making the binaries change from detached to contact and finally coalesce, resulting in a large contribution to BSs in old clusters, e.g NGC 2682. In young and intermediate-age clusters, however, AML has little contribution to the birthrate of BSs, since (a) the time is not long enough for binaries to go from detached to contact and (b) the mass of the mergers is probably less than the turnoff of the cluster even though their parents may coalesce in the cluster age. So, we simply estimated the effect of AML in NGC 2682 while negecting it in NGC 2660.} The mass loss during the merger process can also affect our result, mainly the location on the CMD of the products. As shown in NGC 2660, no BSs have been observed in the region with the most opportunity from our models. Because of mass loss, the mergers will be fainter than those given in the paper. However the faintness will be slight since the mass loss is not vast from both observations and smooth particle hydrodynamic simulations \\cite{lrs96,sill97,sill01}. The lost mass may carry some angular momentum out from the parent binary. By analyzing the BS spectra from Hubble Space Telescope (there is an apparent continuum deficit on the short-wavelength side of Balmer discontinuity ), De Marco et al. \\shortcite{dm04} argued that some BSs might be surrounded by a circumstellar disk. However, Porter \\& Townsend \\shortcite{pt05} showed that the flux deficits may be attributed wholly to rapid rotation. The rotation rates needed are of the order of those found in the study of De Macro et al. \\shortcite{dm05}. Whether the flux deficits shortward of the Balmer jump are induced by a circumstellar disk or rapid rotation, it provides a possible explanation for the orbital angular momentum of the system after coalescence. Such a large mass loss as shown in NGC 2660 (about $0.5 M_\\odot$), however, is a problem and should be explained reasonably in physics. Based on some assumptions, we studied the mergers of close binaries from AS and AR evolution by detailed evolutionary calculations. The products from our models may stay on the left of the ZAMS and have no central concentration with decreasing mass. Because of the {\\bf development} of the convective core, the mergers with little central hydrogen (less than 0.01) in our models have unusually long timescales on the main sequence ($\\sim 10^8$ yrs). These objects are probably subgiants as they are formed, since the primaries in the progenitors also have little central hydrogen and may have left the main sequence during merger process. The mergers from our models stay on the main sequence for a timescale in order of $10^8$ yrs. Some low-mass mergers may stay on the MS for about $10^9$ yrs. The timescale is similar to that of W UMa stars from observations, and therefore we may roughly estimate the contribution to BSs from AS and AR via the number of W UMa systems in a cluster. The estimation, however, is not absolutely since both of the two timescales have wide ranges and large uncertainties, and we cannot rule out other methods for creating W UMa systems except for AS and AR. Comparison to observations indicates that our models (binary coalescence from AS and AR) are not important for the produce of BSs in old open clusters, while likely play a critical role in some younger open clusters. We performed Monte Carlo simulations to examine our models in an old open cluster NGC 2682 and in an intermediate-age cluster NGC 2660. The effect of AML was estimated in NGC 2682 in a simple way, where the mergers are replaced with ZAMS models. In NGC 2682, binary mergers from our models cover the region with high luminosity and those from AML are located in the region with low luminosity, existing a certain width. The BSs from AML are much more than those from our models, indicating that AML of low mass binaries makes a major contribution to BSs in this cluster. Our models are corresponding for several BSs in NGC 2660. In the region with the most opportunity on CMD, however, no BSs have been observed. {\\bf Our results are well-matched to the observations if there are $\\sim 0.5M_\\odot$ of mass loss in the merger process, but a physical mechanism for this much mass loss is a problem.}" }, "0710/0710.5607_arXiv.txt": { "abstract": "Extreme gravitational lensing refers to the bending of photon trajectories that pass very close to supermassive black holes and that cannot be described in the conventional weak deflection limit. A complete analytical description of the whole expected phenomenology has been achieved in the recent years using the strong deflection limit. These progresses and possible directions for new investigations are reviewed in this paper at a basic level. We also discuss the requirements for future facilities aimed at detecting higher order gravitational lensing images generated by the supermassive black hole in the Galactic center. ", "introduction": "All known astrophysical cases of bending of photon trajectories by gravitational fields can be interpreted using the conventional weak deflection paradigm, originally formulated by Einstein \\cite{Ein}. Yet it is clear that photons passing very close to black holes suffer very large deflections, which must be addressed in a full general relativistic context. However, integrating null geodesics in full general relativity usually leads to very complicated analytical formulae or heavy numerical codes, which typically obscure the basic physical interpretation. The Strong Deflection Limit (SDL) is an analytical tool that allows to derive simple analytical formulae describing the higher order images appearing in extreme gravitational lensing. In this work we review the main progresses achieved in the recent years in this technique and its application to the most interesting physical case: the black hole in the Galactic center (Sgr A*). All relevant formulae derived in previous works are recalled and restated here in the simplest possible form. This work is structured as follows: \\S~2 explains the basic phenomenology of extreme lensing; \\S~3 introduces the SDL for spherically symmetric black holes; \\S~4 generalizes the method to spinning black holes; \\S~5 applies the formulae to Sgr A*, examining possible sources for extreme gravitational lensing; \\S~6 generalizes the method to sources very close to the black hole; \\S~7 contains the conclusions. ", "conclusions": "Extreme gravitational lensing is a very spectacular though elusive phenomenon, which demands a great effort in order to be observed. The most promising extreme lens is represented by the supermassive black hole in the Galactic center, which anyway requires resolutions of order microarcseconds for the direct observation of higher order images. Suitable known sources for gravitational lensing are giant stars in the IR band and low mass X-ray binaries in the X-ray band. In this paper we have reviewed recent results on gravitational lensing in the Strong Deflection Limit, which is an approximation devoted to the analytical determination of the properties of higher order images. Within this framework, it has been proved that the logarithmic divergence in the deflection angle is a universal feature of all black hole metrics. There exists a general method to determine the SDL coefficients of the deflection angle for any given spherically symmetric black hole metric. This method has been applied to numerous metrics. Higher order images are formed by photons performing one or more loops around the black hole before reaching the observer. The position and the flux ratios of higher order images depend on the specific metric through the SDL coefficients and are thus able to track any possible deviations from general relativity in the strong field regime. The SDL can be extended to spinning black holes, where several new features emerge, such as extended and shifted caustics and additional images. The position and the shape of the caustics can be expressed by simple analytical formulae and the lens equation for sources near to caustics, which represent the most physically relevant situation, can be solved. The extension of the SDL to the case of sources very close to black holes opens the way to even more interesting investigations of extreme gravitational lensing. In fact, whereas direct observation of higher order images is very difficult and probably very far to come in the future, the contribution of higher order images to currently observed phenomena involving sources very close to black holes is already measurable at present time. Light curves of flares born in the accretion disk and spectral measurements of Iron K-lines are heavily influenced by the presence of higher order images. With the SDL setup in its updated version, it is possible to study such phenomena, bearing in mind that the large model dependence remains the main uncertainty in all theoretical attempts to interpret the complicated physics of the supermassive black holes environment." }, "0710/0710.5431_arXiv.txt": { "abstract": "Warm dark matter is consistent with the observations of the large-scale structure, and it can also explain the cored density profiles on smaller scales. However, it has been argued that warm dark matter could delay the star formation. This does not happen if warm dark matter is made up of keV sterile neutrinos, which can decay into X-ray photons and active neutrinos. The X-ray photons have a catalytic effect on the formation of molecular hydrogen, the essential cooling ingredient in the primordial gas. In all the cases we have examined, the overall effect of sterile dark matter is to facilitate the cooling of the gas and to reduce the minimal mass of the halo prone to collapse. We find that the X-rays from the decay of keV sterile neutrinos facilitate the collapse of the gas clouds and the subsequent star formation at high redshift. ", "introduction": "Both cold and warm dark matter models agree with the observed structure on the large scales. However, there are several inconsistencies between the predictions of the cold dark matter (CDM) model and the observations~\\cite{cdm_problems}. The low cutoff in dark matter contents of dwarf spheroids, the smoothness of our dark matter halo, and the old globular clusters (observed in Fornax) resisting the infall into the center by dynamical friction~\\cite{cdm_problems}, all can be explained by warm dark matter (WDM) because it suppresses the structure on scales that are smaller than the free-streaming length. While the suppression of the small-scale structure is desirable, it has been argued that ``generic'' WDM (for example, gravitino) can slow down structure formation and delay reionization of the universe, which can lead, in turn, to an inconsistency with the reionization redshift obtained by the WMAP \\cite{Yoshida}. This problem can be alleviated in the case of the WDM in the form of sterile neutrinos with mass of several keV and a small mixing angle with the ordinary neutrino \\cite{bier,stas,stas07} because such sterile neutrinos can decay and produce photons that catalyze the formation molecular hydrogen and speed up the star formation. In the absence of metals, gas cooling is mainly due to the collisional excitation of H$_{2}$, its subsequent spontaneous de-excitation, and photon emission. In the primordial gas clouds, hydrogen molecules can be formed only in reactions involving $e^{-}$ or H$^+$ as a catalyst. Thus, an X-ray radiation can increase the production of the H$_{2}$ by enhancing the ionization fraction, which subsequently leads to speed up of the gas cooling and star formation. Although sterile neutrinos are stable on cosmological time scales, they nevertheless decay. The decay channel important for us is that of decay into one active neutrino and one photon, i.e., $\\nu_s \\rightarrow \\nu_a \\gamma$, where the photon energy is half of the sterile neutrino mass, $E_{0} \\approx m_{s}c^2/2$. These decays produce an X-ray background radiation that increases the production of molecular hydrogen and can induce a rapid and prompt star formation at high redshift. Sterile dark matter has a firm motivation from particle physics~\\cite{dw,Kusenko:2006rh}. The discovery of the neutrino masses implies the existence of right-handed gauge-singlet fields, all or some of which can be lighter than the electroweak scale. These sterile neutrinos can be produced in the early universe by different mechanisms, for example, from neutrino oscillations~\\cite{dw} or from the Higgs decays~\\cite{Kusenko:2006rh}, or from the couplings to a low-scale inflaton~\\cite{Shaposhnikov:2006xi}. The same particles, produced in a supernova, could account for the supernova asymmetries and the pulsar kicks~\\cite{kus}, and can play a role in the formation of super-massive black holes in the early universe~\\cite{puzzle}. We will examine the thermal evolution of the gas clouds, taking into account both effects of the sterile neutrino decays, namely, the ionization and heating of the gas. We follow the evolution of the baryonic top-hat overdensity, the gas temperature and the H$_{2}$ and $e^{-}$ fraction. In order to perform the calculation we have incorporated to our previous code \\cite{stas}, the effects of sterile neutrino decays within collapsing halos and absorption of the X-ray background from sterile neutrinos by He atoms in the intergalactic medium. Our goal is to juxtapose the evolution of the gas temperature in the primordial clouds in the CDM model and the WDM model with keV sterile neutrinos and estimate of the minimal mass able to collapse at a given redshift. ", "conclusions": "" }, "0710/0710.2027_arXiv.txt": { "abstract": "{ The exciting results from H.E.S.S. point to a new population of $\\gamma$-ray sources at energies E$>$10~TeV, paving the way for future studies and new discoveries in the multi-TeV energy range. Connected with these energies is the search for sources of PeV cosmic-rays (CRs) and the study of multi-TeV $\\gamma$-ray production in a growing number of astrophysical environments. {\\em TenTen} is a proposed stereoscopic array (with a suggested site in Australia) of modest-sized (10 to 30m$^2$) Cherenkov imaging telescopes with a wide field of view (8$^\\circ$ to 10$^\\circ$ diameter) optimised for the E$\\sim$10 to 100~TeV range. {\\em TenTen} will achieve an effective area of $\\sim$10~km$^2$ at energies above 10~TeV. We outline here the motivation for {\\em TenTen} and summarise key performance parameters. } \\email{growell@physics.adelaide.edu.au} \\begin{document} ", "introduction": " ", "conclusions": "We have outlined the motivation for a new array of IACTs achieving 10~km$^2$ at $E>10$~TeV and described some important performance parameters. This array, known as {\\em TenTen}, could also be considered complementary to future MeV to TeV $\\gamma$-ray instruments such as GLAST, HESS-II and MAGIC-II. Studies are currently underway to further optimise individual telescopes (optics, electronics, camera design), overall layout parameters, and site potential in Australia. \\scriptsize" }, "0710/0710.5788_arXiv.txt": { "abstract": "The observed increase in star formation efficiency with average cloud density, from several percent in whole giant molecular clouds to $\\sim30$\\% or more in cluster-forming cores, can be understood as the result of hierarchical cloud structure if there is a characteristic density as which individual stars become well defined. Also in this case, the efficiency of star formation increases with the dispersion of the density probability distribution function (pdf). Models with log-normal pdf's illustrate these effects. The difference between star formation in bound clusters and star formation in loose groupings is attributed to a difference in cloud pressure, with higher pressures forming more tightly bound clusters. This correlation accounts for the observed increase in clustering fraction with star formation rate and with the observation of Scaled OB Associations in low pressure environments. ``Faint fuzzie'' star clusters, which are bound but have low densities, can form in regions with high Mach numbers and low background tidal forces. The proposal by Burkert, Brodie \\& Larsen (2005) that faint fuzzies form at large radii in galactic collisional rings, satisfies these constraints. ", "introduction": "\\label{sect:intro} Stars form in concentrations with a range of densities, from star complexes, OB associations, and T Tauri associations at the low end to compact clusters and super-star clusters (SSC) at the high end. The overall structure is usually hierarchical (Scalo 1985; Feitzinger \\& Galinski 1987; Ivanov et al. 1992; Gomez et al. 1993; Battinelli, Efremov \\& Magnier 1996; Bastian, et al. 2005, 2007; Elmegreen et al. 2006; see reviews in Efremov 1995; Elmegreen et al. 2000, Elmegreen 2005), and this hierarchy continues even inside the youngest clusters (Testi et al. 2000; Heydari-Malayeri et al. 2001; Nanda Kumar, Kamath, \\& Davis 2004; Smith et al. 2005a; Gutermuth et al. 2005; Dahm \\& Simon 2005; Stanke, et al. 2006; see review in Allen et al. 2006). Most likely, the hierarchy in stars comes from a hierarchy in the gas (St\\\"utzki et al. 1998; Dickey et al. 2001), which is the result of turbulence compression and gravitational contraction that is self-similar over a wide range of scales (see review in Mac Low \\& Klessen 2004). Clusters form in the densest parts of this gas and lose their initial sub-structure as the stellar orbits mix (for simulations of star formation in clusters, see Klessen \\& Burkert 2000; Bonnell, Bate, \\& Vine 2003; Li, et al. 2004; Tilley \\& Pudritz 2004; Klessen et al. 2005; V\\'azquez-Semadeni, Kim, \\& Ballesteros-Paredes 2005; Bonnell \\& Bate 2006; Li \\& Nakamura 2006). Hunter (1999) and Ma\\'iz-Apell\\'aniz (2001) noted that some massive star-forming regions (which they called ``scaled OB associations'', or SOBA's) do not form dense clusters while others with the same total mass do (e.g., the SSC's). We would like to understand this difference. Obviously the density of the gas is involved, as dense clusters require dense gas, but the distinction between SSCs and SOBAs should also be related to the efficiency of star formation, because clusters forming at low efficiency disperse quickly when the gas leaves (Lada, Margulis \\& Dearborn 1984). At very low efficiency, stars form individually without passing through an embedded cluster phase. Recent observations of giant molecular clouds (GMCs) show star formation at both high and low densities, with some embedded stars in dense clusters and others more dispersed (Megeath et al. 2004; J{\\o}rgensen et al. 2006, 2007). Larsen \\& Brodie (2000) discovered ``faint fuzzies'' at intermediate radii in the disks of the S0 galaxies NGC 1023 and NGC 3384, and suggested they are old clusters with unusually large radii (7-15 pc) and low densities. They are gravitationally bound because of their large ages (8-13 Gyr; Brodie \\& Larsen 2002), and they are as massive as SSC's and halo globular clusters. Such clusters appear to represent an intermediate stage between dispersed and bound star formation. They appear to be too low in average density to have had time for core collapse and envelope expansion as in models of globular clusters by Baumgardt et al. (2002). Burkert, Brodie \\& Larsen (2005) suggested they formed in a collisional ring interaction between two galaxies. What determines the relative proportion of dispersed and clustered star formation? Larsen \\& Richtler (2000) showed that clustering on a galactic scale, measured as the fraction of uv light in the form of massive young clusters, increases as the star formation rate increases. This could be the result of a selection effect if starbursts are active for less than a cluster destruction time. On the other hand, the clustering fraction could depend on pressure. Higher pressure makes the cool phase of gas denser, which promotes more clustering, and locally high pressures trigger star formation on the periphery of GMCs, making clusters in the dense gas (e.g., Comer\\'on, Schneider, \\& Russeil 2005; Zavagno et al. 2006). The Larsen \\& Richtler correlation could then follow from the mutual correlation between pressure and star formation rate with gas column density. Faint fuzzies are a counter example, however: they formed bound but presumably at low pressure to have such low central column densities now. Can all of these clustering types be understood as a continuum of properties in a universal physical model? There have been several attempts to explain the difference between clustered and distributed star formation based on numerical simulations. Klessen, Heitsch, \\& Mac Low (2000) suggested that clusters form in non-magnetic gas when the turbulence driving scale is large. Heitsch, Mac Low \\& Klessen (2001) noted that non-magnetic turbulence driven on large scales produces a clustered collapse, while magnetic turbulence in supercritical clouds produces a more distributed collapse. Mac Low (2002) suggested that stars form in clusters when there is no turbulent support and they form disbursed when there is. V\\'azquez-Semadeni, Ballesteros-Paredes, \\& Klessen (2003) suggest this transition from no global turbulent support to support corresponds to an increase in the Mach number and a decrease in the sonic scale, which is the length where the size-linewidth relation gives a Mach number of unity. Large sonic scale (low Mach number) corresponds to a sonic mass larger than the thermal Jeans mass, which means a lack of global support and the formation of a cluster. Low sonic scale corresponds to the dispersed formation of stars, one for each tiny compressed region where the mass exceeds the local thermal Jeans mass. Li, Klessen \\& Mac Low (2003) suggested that the equation of state determines the stellar clustering properties: soft equations produce dense clusters while hard equations produce isolated stars. These suggestions all apply to initially uniform media. External compression of a cloud into a massive dense core can also make a cluster; most young clusters are in high-pressure regions like OB associations. Simulations of turbulent media produce stars in compressed regions that act as seeds for the small-scale gravitational collapses that follow (V\\'azquez-Semadeni, Ballesteros-Paredes, \\& Klessen 2003; Clark \\& Bonnell 2005). These simulations also have probability distribution functions (pdfs) for density that are either log-normal or log-normal with a power-law tail at high density, especially when self-gravity is important (e.g., Li, Klessen \\& Mac Low 2003). The efficiency of star formation is then proportional to the fraction of the gas in a dense form. Here we examine variations in this fraction as functions of average density and velocity dispersion, and as a function of the local density inside a cloud. The efficiency is taken to be the ratio of the stellar mass to the gas mass during a complete star-forming event. It generally increases with cloud density from a few percent in GMCs (Williams \\& McKee 1997) to several tens of percent in cluster-forming cores (e.g., Lada \\& Lada 2003). It may reach $\\sim50$\\% or more inside the densest star-forming cores. We explain this increase as a result of hierarchical structure, regardless of the dynamics and mechanisms of star formation, and we show that for log-normal or similar density pdf's, as expected in turbulent media, the mass fraction of regions with high efficiency increases with the Mach number and, independently, with the average density. This result may explain the Larsen \\& Richtler (2000) correlation as well as the observed variations in clustering properties with pressure. We also show that at high Mach number, bound clusters can form with relatively low average densities, thereby explaining faint fuzzies. These are all consequences of star formation in the dense cores of clouds that are structured by turbulence. They result primarily from the geometry of the gas, which is somewhat universal, and should be nearly independent of the gas dynamics or the strength of the magnetic field. We make an important assumption that gravitational contraction and star formation can occur in regions that are either larger or smaller than the sonic scale. This means we assume that contraction to one or more stars can occur in a supersonically turbulent region. V\\'azquez-Semadeni, Ballesteros-Paredes, \\& Klessen (2003) suggest that if a cloud is supported by turbulence, then only regions smaller than a sonic scale and more massive than the thermal Jeans mass are unstable to form stars. Padoan (1995) was the first to consider this condition. However, clouds are probably not supported for any significant time by turbulence, and even if they were, it is only necessary that a clump mass exceed the turbulent Jeans mass for self-gravitational forces to exceed inertial forces. GMCs for example, have comparable self-gravitational and turbulent energy densities and yet are much larger than the sonic length. Our assumption is contrary to that in Krumholz \\& McKee (2005), who assume the same as V\\'azquez-Semadeni et al.. We are consistent with McKee \\& Tan (2003), however, as they consider the collapse of a highly turbulent core to make a massive star. Saito, et al. (2006), for example, observe star formation in massive turbulent cores. Thus the sonic length should not provide a threshold for star formation. Our primary condition for star formation is that the gas density exceed some fixed value, taken here to be $10^5$ cm$^{-3}$. This is the density at which HCN gives a nearly constant star formation rate per unit gas mass (Gao \\& Solomon 2004a,b; Wu et al. 2005) and at which a variety of microscopic processes conspire to shorten the magnetic diffusion time (Elmegreen 2007). Regions with this density should have a wide range of Mach numbers but a nearly universal efficiency, according to the Gao \\& Solomon and Wu et al. observations. The density of $\\sim10^5$ cm$^{-3}$, when converted to 5900 M$_\\odot$ pc$^{-3}$, is also typical for star clusters, as most of those surveyed by Tan (2007), which span a factor of $10^5$ in mass, have about this average density. The cluster density equals the gas density times the efficiency, and the efficiency has to exceed $\\sim10-30$\\% for a bound cluster to form. Thus, the gas density for both cluster formation and high efficiency appears to be around $10^5$ cm$^{-3}$ or, possibly, $10^6$ cm$^{-3}$. We note that the long timescale derived for HCN gas as the ratio of the total HCN mass divided by the total star formation rate (Gao \\& Solomon 2004b; Wu et al. 2005) is not the duration of star formation in any one place, but is the HCN consumption time. As long as a new HCN region is formed somewhere each time an old HCN region disperses, the HCN consumption time can be long even when each region of star formation lasts for a short time (see Elmegreen 2007 for a discussion of star formation timescales). As a result, the average efficiency of star formation in any one HCN region can be moderately large even if the average HCN consumption time is long. An average efficiency of $\\sim5-10$\\% would be reasonable considering that most star-forming regions leave unbound clusters after the gas leaves (Lada \\& Lada 2003), most regions are observed at half their total ages, and star formation typically accelerates over time (Palla \\& Stahler 2000). With constant acceleration, only $\\sim1/4$ of the total stars form in the first 1/2 of the total time. The efficiency that determines whether a bound cluster will remain is taken here to be 14.4\\% (see below), and the peak efficiency in a single-star core is taken to be 50\\% at $10^5$ cm$^{-3}$ density. These values are uncertain and are used here only to illustrate how cluster formation might scale with the velocity dispersion and density of the cool component of the interstellar medium. ", "conclusions": "Stars form bound clusters where the total efficiency is high. The efficiency should be proportional to the mass fraction of a cloud in the form of dense cores where the individual stars form. This mass fraction increases with the cloud density in hierarchical clouds. It also increases with the Mach number of the turbulence because stronger shocks at higher Mach numbers compress the gas to a wider range of densities. As a result, a log-normal density pdf becomes flatter below the threshold density of star formation, and this means the mass fraction is higher at each density below this value. When the Mach number is high (really, when the dispersion of the density pdf is high), the efficiency is high even at a fairly low average cloud density, and so a high fraction of star formation ends up bound. An extreme example of this trend is the type of cluster called a faint fuzzie. We propose that faint fuzzies form in moderately low density clouds with moderately high Mach numbers. The galactic tidal ring environment proposed by Burkert et al. (2005) is an example of a region that would have such clouds. The mass fraction of clumps at a particular high density of star formation also increases with the average density of the ISM because then the whole density pdf shifts toward higher values. The combination of high densities and high Mach numbers, characteristic of starburst regions, makes for a high fraction of star formation in bound clusters. On the other hand, relatively low Mach numbers and/or low densities should produce stars in a more dispersed way, as in the low-density regions of molecular clouds or in low surface brightness galaxies and regions of galaxies. This combination of parameters corresponds to a low ISM pressure, so we infer that low pressure regions, which means those with a low gas column densities and low star formation rates per unit area, should produce relatively fewer bound clusters and relatively more unbound associations. There is a physical explanation for the trends discussed here. Consider a moderately low density cloud with a low turbulent Mach number. The compressions inside that cloud will be modest and most of them will not reach a level of density or enhanced magnetic diffusion rate that allows gravitational collapse before the compression ends. Then few stars will form and the efficiency will be low (unless the cloud continues to makes these weak compressions for a very large number of crossing times, which seems unlikely). Now consider this same cloud with the same size and average density but with a higher Mach number (this will require a higher ISM pressure). The stronger compressions will more easily produce high-density cores in which gravity is important and magnetic diffusion is fast. More stars will form and the efficiency will be high. The only difference between these two examples is the Mach number, which affects the range of densities in the compressed regions. In terms of the density pdf, higher Mach numbers produce a broader pdf at the same average density. In terms of star formation in bound clusters, high pressure regions having clouds with high Mach number produce a higher fraction of their stars in bound clusters. The model predicts that young bound clusters in starburst regions, or in regions of high ISM pressures or Mach numbers, will have a wider range of densities than bound clusters in more quiescent, low-pressure regions. This is partly because the cloud densities should be higher in starburst regions, so the resulting clusters can be denser overall, but it is also because the lower-density parts of starburst clouds can form stars with a high efficiency, leaving bound clusters when the gas leaves rather than dispersed OB associations. As there is generally more mass at low density than at high density, the net distribution of cluster density could shift toward lower values when the ISM density pdf is broad. According to the model, this density shift is relative to the mean ISM density, and it should be measured only before significant cluster expansion. Low surface brightness galaxies should make a preponderance of unbound OB associations rather than bound clusters, and the young clusters they form should have a relatively narrow range of densities compared to normal." }, "0710/0710.0955_arXiv.txt": { "abstract": "{} {We present the optical classification and redshift of 348 X-ray selected sources from the XMM-Newton Bright Serendipitous Survey (XBS) which contains a total of 400 objects (identification level = 87\\%). About 240 are new identifications. In particular, we discuss in detail the classification criteria adopted for the Active Galactic Nuclei population.} {By means of systematic spectroscopic campaigns and through the literature search we have collected an optical spectrum for the large majority of the sources in the XBS survey and applied a well-defined classification ``flow-chart''. } {We find that the AGN represent the most numerous population at the flux limit of the XBS survey ($\\sim$10$^{-13}$ erg cm$^{-2}$ s$^{-1}$) constituting 80\\% of the XBS sources selected in the 0.5-4.5 keV energy band and 95\\% of the ``hard'' (4.5-7.5 keV) selected objects. Galactic sources populate significantly the 0.5-4.5 keV sample (17\\%) and only marginally (3\\%) the 4.5-7.5 keV sample. The remaining sources in both samples are clusters/groups of galaxies and normal galaxies (i.e. probably not powered by an AGN). Furthermore, the percentage of type~2 AGN (i.e. optically absorbed AGNs with A$_V>2$mag) dramatically increases going from the 0.5-4.5 keV sample (f=N$_{AGN 2}$/N$_{AGN}$=7\\%) to the 4.5-7.5 keV sample (f=32\\%). We finally propose two simple diagnostic plots that can be easily used to obtain the spectral classification for relatively low redshift AGNs even if the quality of the spectrum is not good. } {} ", "introduction": "In the last few years {\\it XMM-Newton} and {\\it Chandra} telescopes have represented an excellent tool to survey the hard X-ray sky at all fluxes, from relatively bright (10$^{-13}$ erg cm$^{-2}$ s$^{-1}$, e.g. Della Ceca et al. 2004 and references therein), to medium (10$^{-13}$ erg cm$^{-2}$ s$^{-1}$-10$^{-14}$ erg cm$^{-2}$ s$^{-1}$, e.g. Barcons et al. 2007 and references therein) and deep (10$^{-14}$-10$^{-16}$ erg cm$^{-2}$ s$^{-1}$, Brandt \\& Hasinger 2005; Worsley et al. 2005 and references therein) fluxes. At the energies ($\\sim$0.5-10 keV) covered by the instruments on board these two telescopes, Active Galactic Nuclei (AGN) can be efficiently selected and studied even when affected by large levels of absorption (up to N$_H\\sim$10$^{24}$ cm$^{-2}$, corresponding to an optical absorption of A$_V\\sim$500 mag). This important characteristic, combined with the good/excellent spatial and energy resolutions of the detectors, makes the ongoing surveys a fundamental tool for AGN studies. At the same time, these new surveys represent an observational challenge at wavelengths different from the X-ray ones: multiwavelength follow-ups of X-ray sources, particularly in the optical domain, are decisive to derive the distance and to understand the properties of the selected objects but they also require large fractions of dedicated observing time at different telescopes. Probably one of the most challenging and time-consuming efforts is the optical spectroscopic follow-up of the selected sources. One of the primary goals of all these hard X-ray surveys is to explore the population of absorbed AGN and, to this end, an optical classification that can reliably separate between optically absorbed and non-absorbed objects is always required. Two important limits, however, affect the spectroscopic follow-ups of deep and, in part, medium surveys: first, the optical counterparts are often too faint to be spectroscopically observed even at the largest optical telescopes currently available; second, even when a spectrum can be obtained, its quality is not always good enough to provide the critical pieces of information that are required to assess a reliable optical classification. These two problems often limit the final scientific results that are based on the optical classification of medium/deep surveys. On the contrary, bright surveys offer the important possibility of obtaining a reliable optical classification for virtually all (with some exceptions, as discussed in the next sections) the selected sources. The disadvantage of dealing with shallow/wide-angle samples is that the techniques to observe efficiently many sources at once, like Multi-objects or fibers-based methods, cannot be applied for the optical follow-up, given the low space-density of sources at bright X-ray fluxes. The only suitable method, the ``standard'' long-slit technique, requires many independent observing nights to achieve the completion of the optical follow-up. In this paper we present and discuss in detail the optical classification process of the {\\it XMM-Newton} Bright Serendipitous Survey (XBS, Della Ceca et al. 2004), which currently represents the widest (in terms of sky coverage) among the existing {\\it XMM-Newton} or {\\it Chandra} surveys for which a spectroscopic follow-up has been almost completed. The aim of the paper, in particular, is to provide not only a generic classification of the sources and their redshift but also a quantification, in the limits of the available data, of the corresponding threshold in terms of level of optical absorption. The paper is organized as follows: in Section~2 we describe the XBS survey, in Section~3 we describe the process of identification of the optical counterpart, in Section~4 and Section~5 we respectively summarize our own spectroscopic campaigns carried out to collect the data as well as the data obtained from the literature. In Section~6 we shortly discuss the data reduction and analysis of the optical spectra and in Section~7 we give the details on the classification criteria adopted for the sources in the XBS survey. In Section~8 we propose two diagnostic plots that can be used to easily classify the sources into type~1 and type~2 AGN. The resulting catalogue is presented in Section~9 while in Section~10 we briefly discuss the optical breakdown and the redshift distribution of the sources. The conclusions are finally summarized in Section~11 . Throughout this paper H$_0$=65 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\Lambda}$=0.7 and $\\Omega_M$ = 0.3 are assumed. ", "conclusions": "We have presented the details of the identification work of the sources in the XBS survey, which is composed by two complete flux limited samples, the BSS and the HBSS sample, selected in the 0.5-4.5 keV and 4.5-7.5 keV band respectively. We have secured a redshift and a spectroscopic classification for 348 (including data from the literature) out of 400 sources, corresponding to 87\\% of the total list of sources and to 87\\% and 97\\% considering the BSS and HBSS samples separately. The results of the identification work can be summarized as follows: \\begin{itemize} \\item We have quantified the criteria used to distinguish optically absorbed AGN (i.e. type~2) from optically non-absorbed (or moderately absorbed) AGN (type~1) and we have shown that the adopted dividing line between the two classes of sources corresponds to an optical extinction of A$_V\\sim$2 mag, which translates into an expected column density of N$_H\\sim$4$\\times$10$^{21}$ cm$^{-2}$, assuming a Galactic A$_V$/N$_H$ relationship. \\item About 10\\% of the extragalactic sources (35 objects in total) show an optical spectrum which is highly contaminated by the star-light from the host galaxy. These sources have been studied in detail in a companion paper (Caccianiga et al. 2007). Using the X-ray data we have found an elusive AGN in 33 of these objects and we have classified them into type~1 and type~2 AGN according to the value of N$_H$ measured from the X-ray spectrum. To this end, we have used a N$_H$=4$\\times$10$^{21}$ cm$^{-2}$ dividing value which matches (assuming the standard Galactic A$_V$/N$_H$ relation) the value of A$_V$ (=2 mag) adopted with the optical classification. \\item We have then proposed two simple diagnostic diagrams. The first one, based on the 4000\\AA\\ break and the [OIII]$\\lambda$5007\\AA\\ equivalent width, can reliably distinguish between type~1 and type~2 AGN if the host galaxy does not dominate the optical spectrum. The second uses the H$\\alpha$ and [OIII]$\\lambda$5007\\AA\\ line equivalent widths to classify into type~1 and type~2 the elusive AGN sources in which a possibly broad H$\\alpha$ emission line is detected. \\item We find that the AGN represent the most numerous population at the flux limit of the XBS survey ($\\sim$10$^{-13}$ erg cm$^{-2}$ s$^{-1}$) constituting 80\\% of the XBS sources selected in the 0.5-4.5 keV energy band and 95\\% of the ``hard'' (4.5-7.5 keV) selected objects. Galactic sources populate significantly the 0.5-4.5 keV sample (17\\%) and only marginally (3\\%) the 4.5-7.5 keV sample. The remaining sources in both samples are clusters/groups of galaxies and normal galaxies (i.e. probably not powered by an AGN). \\item As expected, the percentage of type~2 AGN dramatically increases going from the 0.5-4.5 keV sample (f=N$_{AGN 2}$/N$_{AGN}$=7\\%) to the 4.5-7.5 keV sample (f=32\\%). A detailed analysis on the intrinsic (i.e. taking into account the selection effects) relative fraction of type~1 and type~2 AGN will be be presented in a forthcoming paper (Della Ceca et al. 2007, in prep.). \\end{itemize}" }, "0710/0710.2175_arXiv.txt": { "abstract": "Voids are a dominant feature of the low-redshift galaxy distribution. Several recent surveys have found evidence for the existence of large-scale structure at high redshifts as well. We present analytic estimates of galaxy void sizes at redshifts $z \\sim 5 - 10$ using the excursion set formalism. We find that recent narrow-band surveys at $z \\sim 5 - 6.5$ should find voids with characteristic scales of roughly $20$ comoving $\\Mpc$ and maximum diameters approaching $40 \\Mpc$. This is consistent with existing surveys, but a precise comparison is difficult because of the relatively small volumes probed so far. At $z \\sim 7 -10$, we expect characteristic void scales of $\\sim 14 - 20$ comoving $\\Mpc$ assuming that all galaxies within dark matter haloes more massive than $10^{10} \\Msol$ are observable. We find that these characteristic scales are similar to the sizes of empty regions resulting from purely random fluctuations in the galaxy counts. As a result, true large-scale structure will be difficult to observe at $z \\sim 7 -10$, unless galaxies in haloes with masses $\\la 10^9 \\Msol$ are visible. Galaxy surveys must be deep and only the largest voids will provide meaningful information. Our model provides a convenient picture for estimating the ``worst-case'' effects of cosmic variance on high-redshift galaxy surveys with limited volumes. ", "introduction": "The complex network of filaments and voids observed in the present-day Universe is believed to have formed from an initially homogeneous distribution of matter. In hierarchical models of structure formation, tiny perturbations seeded by the inflationary epoch grew through gravitational instability, collapsing first on smaller scales to form haloes. The subsequent merging and clustering of smaller haloes resulted in the formation of highly structured large-scale systems. Perhaps the most striking characteristic of the Universe today is the prevalence of large and nearly spherical voids in the galaxy distribution. The scales of these voids can be enormous. Indeed, \\citet{HandV2004} report characteristic radii of $R \\sim 15h^{-1}~\\Mpc$ with maximum scales approaching $R \\sim 25 h^{-1} \\Mpc$ in the 2dF Galaxy Redshift Survey. The characteristics of voids and the galaxies that populate them have been the subject of numerous theoretical and observational studies throughout the years \\citep{gregory78, kirshner81, delapparent86, vogeley94, HandV2004, Conroy2005}. To date, these studies have mostly focused on low redshifts. However, it is clear that voids should appear at higher redshifts also. The DEEP2 survey indicates that voids exist at redshifts of $z \\sim 1$ \\citep{Conroy2005}. Surprisingly, a handful of recent Lyman $\\alpha$ emitter (LAE) surveys have found hints of large-scale structure at redshifts around $z \\sim 5$ \\citep{Shimasaku2003, Shimasaku2006, Hu2004, Ouchi2005}. The modelling of voids poses an interesting theoretical problem. There have been numerous studies utilising $N$-body simulations \\citep{MandW2002,Benson2003,Gottlober2003,Colberg2005}. While these simulations are invaluable tools for understanding the details of void dynamics, they are computationally expensive due to the large volumes and high dynamic range required to include a representative sample of voids while also resolving the much smaller galaxies that define them. Analytic methods provide a useful alternative. Perhaps the most promising analytic model of void abundances is the excursion set approach taken by \\citet{SandV2004}. They argue that voids actually provide deeper insight into large-scale structure than halo formation itself. Their assertion is based on the fact that underdense regions generally tend to evolve toward a spherical geometry, making the idealisation of spherical expansion more reasonable. In contrast, gravitationally bound objects typically have geometries that are far from spherical. Approximating gravitational collapse with the spherical model may be highly inaccurate, which partially explains the discrepancies between the \\citet{PandS1974} halo mass function and simulations \\citep{SandT1999,Sheth2001,Jenkins2001}. A key disadvantage to the approach of \\citet{SandV2004} is the difficulty in relating their definition of voids to observational studies. As we will discuss in \\S \\ref{VoidDef}, \\citet{SandV2004} use the dark matter underdensity to define voids. \\citet{FandP2006} extend their model to define voids in terms of the local \\emph{galaxy} underdensity. They predict characteristic void sizes of $R \\sim 10 h^{-1} \\Mpc$ at the present day -- nearly as large as observed voids. In what follows, we present analytic estimates of void size distributions at redshifts between $z \\sim 5 - 10$. Our aim is to provide a convenient basis of comparison for current and future high-redshift observations -- presumably, though not limited to, LAE surveys. As such, we consider the effects of statistical fluctuations in the galaxy counts and the abundance of Ly $\\alpha$ emitting galaxies on void observations. LAE surveys have become an invaluable tool in cosmological studies. In addition to building larger samples at $z \\sim 5$, observers have pushed the threshold to redshifts as high as $z \\sim 7 - 10$ \\citep{WandC2005,Cuby2007,Stark2007,Ota2007}. Surveys at these redshifts could potentially reveal important information on the epoch of reionization. Indeed, the observed clustering properties of LAEs could someday be a powerful probe of the epoch \\citep{furl04-lya, Furlanetto2006,McQuinn2006,McQuinn2007, mesinger07}. Regions of ionized hydrogen grow quickly around clustered galaxies as reionization progresses. When these regions are large enough, Ly $\\alpha$ photons are sufficiently redshifted before they reach neutral hydrogen gas, allowing them to avoid absorption in the intergalactic medium (IGM). Sources within overdense regions are therefore more likely to be observed relative to void galaxies, resulting in a large-scale modulation of the number density. One method to quantify such clustering is with void statistics, as first attempted by \\citet{McQuinn2007}. This provides comparable power to correlation function measurements of the galaxies. However, taking full advantage of this technique requires a deeper understanding of voids in the underlying galaxy distribution; our calculations aim to provide such a baseline model. The remainder of this paper is organized in the following manner. In \\S \\ref{ExcursionSet}, we briefly review the basic principles of the excursion set formalism. In section \\ref{VoidDef}, we present the \\citet{FandP2006} definition of voids in terms of the local galaxy underdensity. Section \\ref{VoidDistributions} contains the main results of this paper: void size distributions at $z = 4.86 - 10$. In \\S \\ref{StochasticVoids}, we estimate the typical sizes of voids that result from random fluctuations in the galaxy distribution and develop an alternative definition of voids. In \\S \\ref{VisibleFrac}, we explore the assumption that only a certain fraction of galaxies are actually visible in LAE surveys. Section \\ref{Observations} contains a rough comparison of our calculations to high redshift Ly $\\alpha$ surveys. Finally, we offer concluding remarks in \\S \\ref{Discussion}. In what follows, we assume a cosmology with parameters $\\Omega_m = 0.24,~\\Omega_{\\Lambda} = 0.76,~\\Omega_b = 0.042,~H = 100h~\\mathrm{km~s}^{-1}~\\Mpc^{-1}$ (with $h = 0.73$), $n = 0.96$, and $\\sigma_8 = 0.8$, consistent with the latest measurements \\citep{Spergel2007}. All distances are reported in comoving units. ", "conclusions": "\\label{Discussion} We have calculated void size distributions at $z = 4.86$--$10$ using the excursion set model developed by \\citet{SandV2004} and \\citet{FandP2006}. The latter found characteristic void radii of $R \\approx 7-14 \\Mpc$ at $z = 0$. For the observational sensitivities assumed in this paper, we obtained characteristic void radii that are very similar: $R \\approx 7-10 \\Mpc$ for redshifts between $z = 4.86$ and $z = 10$. These results are virtually independent of the void-crushing barrier (for any reasonable choice). We have shown that characteristic void scales actually increase with redshift for a fixed halo mass threshold due to a decreased number density and increased bias with respect to the underlying matter density. Following recent studies on the abundances of low-redshift LAEs, we explored the possibility that only a fraction $f_{vis}$ of galaxies are sampled in LAE surveys. This has only a small effect on the void size distribution but increases the contamination of void samples by stochastic fluctuations. In section \\ref{StochasticVoids}, we have explored stochastic fluctuations in the galaxy distribution. These fluctuations, although inherently different from the \"real\" voids we model in this paper, will result in large empty regions in the sky. Stochastic voids can therefore contaminate real void samples and lead to erroneous conclusions on the formation of large-scale structure. We have estimated the typical scale of these regions to be slightly smaller than the characteristic scale of true voids at $z \\sim 5$. At $z \\sim 10$, the situation depends on the particular choice of $m_{min}$. For $m_{min} \\sim 10^{10} \\Msol$, stochastic voids are typically the same scale as real voids. The increased importance of stochastic fluctuations will make the identification of large-scale structure at this redshift difficult. Attempts to do so must observe halos near the minimum mass to form stars, $\\sim 10^8$--$10^9 \\Msol$, in order for true voids to dominate the observed distribution. We found that a large fraction of real voids in our fiducial model contain no visible galaxies, adding to the difficulties in differentiating them from stochastic fluctuations. We have presented a modified definition of voids that incorporates both stochastic and real voids and so is easier to compare to the limited observational samples thus far available. In our new approach, we defined voids in terms of the probability for a region to be empty. We found that the modified void distributions are more sharply peaked and have characteristic scales that are comparable to the fiducial model. We have also attempted to visually compare our results to the most recent narrow-band filter surveys at $z = 5.7$. While we found no inconsistencies, it is difficult to draw any decisive conclusions because of small-number statistics, projection effects, and lower-redshift contaminants. Obviously, a more systematic approach is required. Future surveys promise to provide better statistics and increased sample volumes for studies on high-redshift voids. In the context of next generation surveys for high-redshift galaxies, our model is useful for gauging the impact of cosmic variance. Consider a fictitious survey at $z = 10$ with a detection threshold of $m_{min} = 10^{10} \\Msol$. Figure \\ref{PoissonPlot} illustrates that stochastic voids with $R \\sim 10 - 20 \\Mpc$ will dominate the sky distribution. Therefore, one must either search for voids with $R > 20 \\Mpc$ or search deeper for significantly smaller sources. The latter may be possible if the sources observed by \\citet{Stark2007} are indeed at $z \\sim 9$, in which case they imply that halos near $\\la 10^9 \\Msol$ are visible \\citep{mesinger07}. However, high-redshift galaxies are so highly biased that even with deep observations, a substantial fraction of the Universe is filled with empty or nearly-empty regions. For example, at $z=10$ and $m_{min} = 10^{10} \\Msol$, $\\sim 37\\%$ of space is filled by regions that are at least 80\\% underdense in galaxies and at least 20 Mpc across -- or fully 7 arcmin. With the small fields of view available to near-infrared detectors, this suggests that either many independent fields must be observed or a large contiguous volume surveyed to be guaranteed of detecting a reasonable number of sources. Finally, we have neglected reionization and its effect on the appearance of large-scale structure. Regions of neutral hydrogen are expected to modulate the LAE density on large scales and accentuate the appearance of structure \\citep{furl04-lya, Furlanetto2006,McQuinn2006,McQuinn2007, mesinger07}. Although the precise time frame is currently unknown, quasar observations and cosmic microwave background measurements have provided some evidence that reionization occurred between $z \\sim 6 - 10$ (e.g, \\citealt{Fan2006,Page2007,MandH2004,MandH2007}). Interestingly, \\citet{Kashikawa2006} found a significant high-luminosity suppression in the LAE luminosity function between $z = 5.7$ and $z = 6.5$. Whether or not reioinization is responsible for this effect is currently unclear \\citep[no such suppression was observed by][]{dawson07}. Because IGM absorption modulates the LAE density on large scales, we would expect reionization to have a substantial effect on the observed void sizes in such narrow-band surveys (it should not affect galaxies identified through broadband effects). Of course, the plots in Figure \\ref{Z7to10} provide analytic estimates only of the \\emph{intrinsic} void size distributions. They provide a basis for comparison with high-redshift surveys in order to determine whether the observed features are easily attributable to the large-scale clustering alone. It therefore helps illuminate efforts to use voids to constrain the IGM properties during reionization, as first attempted by \\citet{McQuinn2007}." }, "0710/0710.3345_arXiv.txt": { "abstract": "Determining the final spin of a black-hole (BH) binary is a question of key importance in astrophysics% . Modelling this quantity in general is made difficult by the fact that it depends on the 7-dimensional space of parameters characterizing the two initial black holes. However, in special cases, when symmetries can be exploited, the description can become % simpler. For black-hole binaries with unequal masses but with equal spins which are aligned with the orbital angular momentum, we show that the use of recent simulations and basic but exact constraints derived from the extreme mass-ratio limit allow to model this quantity with a simple analytic expression. Despite the simple dependence, the expression models very accurately all of the available estimates, with errors of a couple of percent at most. We also discuss how to use the fit % to predict when a Schwarzschild BH is produced by the merger of two spinning BHs, when the total angular momentum of the spacetime ``flips'' sign, or under what conditions the final BH is ``spun-up'' by the merger. % Finally, suggest an extension of the fit % to include unequal-spin binaries, thus potentially providing a \\textit{complete} description of the final spin from the coalescence of generic black-hole binaries with spins aligned to the orbital angular momentum. ", "introduction": "\\label{intro} The determination of the final spin of a BH binary is a question of key importance in astrophysics. Modelling this in general is made difficult by the fact that it depends on the 7-dimensional space of parameters characterizing the two initial BHs. However, in special cases, when symmetries can be exploited, the description can be much simpler. Several recent studies have shed light on the remnant of the merger process. Using conservation principles, Hughes and Blandford~\\citep{Hughes_Blandford:2003} argued that mergers rarely lead to rapidly rotating objects. \\citet{Gonzalez:2006md} numerically evolved a sequence of non-spinning unequal-mass BHs, arriving at detailed estimates of the radiated energy and angular momentum. In a series of papers~\\citep{Koppitz:2007ev, Pollney:2007ss, Rezzolla_etal:2007a} we have studied the parameter space of mergers of equal-mass BH binaries whose spins are aligned with the orbital angular momentum but otherwise arbitrary. The findings agree well with independent numerical evolutions~\\citep{Campanelli_etal:2007,Herrmann:2007ac}, as well as more recent studies of models with initial spins up to $J/M^2=0.8$~\\citep{Marronetti_etal:2007}. An important result of these studies has been the determination of simple (quadratic) fitting formulas for the recoil velocity and spin of the merger remnant as a function of the initial BH parameters~\\citep{Rezzolla_etal:2007a}. A number of analytical approaches have been developed over the years to determine the final spin of a binary coalescence~\\citep{ Damour:2001tu, Buonanno_Damour:2000, Buonanno_etal:2006, Damour_Nagar:2007a, Boyle:2007sz}. Very recently, an interesting method, inspired by the dynamics of a test particle around a Kerr BH, has been proposed for generic binaries~(\\citet{Buonanno_etal:2007b}, BKL hereafter). The approach assumes that the angular momentum of the final BH is the sum of the individual spins and of the orbital angular momentum of a test particle on the last-stable orbit of a Kerr BH with the same spin parameter as that of the final BH. Here, we combine the data obtained in recent simulations to provide a phenomenological but analytic estimate for the final spin in a binary BH system with arbitrary mass ratio and spin ratio, but in which the spins are constrained to be parallel to the orbital angular momentum. Our numerical simulations have been carried out using the CCATIE code~\\citep{Pollney:2007ss}. In addition to the data presented in \\citet{Rezzolla_etal:2007a}, we add three simulations of equal-mass, high-spin binaries and three simulations of unequal-mass, spinning binaries (see Table~\\ref{tableone}). Other data is taken from % unequal-mass, nonspinning binaries~\\citep{Gonzalez:2006md,Berti_etal:2007,Buonanno_etal:2007a}, and of equal-mass, spinning binaries~\\citep{Rezzolla_etal:2007a,Marronetti_etal:2007}; all of the AEI data is summarized in Table~\\ref{tableone}. To avoid the possible contamination from the errors associated with high-spin binaries reported by~\\citet{Marronetti_etal:2007}, we have not considered binaries with initial spin $|J/M^2| \\geq 0.75$ reported in the literature~\\citep{Campanelli_etal:2007,Marronetti_etal:2007}. We have, however, considered estimates of high-spin binaries (\\textit{cf.}, Table~\\ref{tableone}), for which we know the spins remain essentially constant prior to merger, with changes less than $0.5\\%$~\\citep{Pollney:2007ss}, and that are very well captured by the fit. ", "conclusions": "Modelling the final spin in a generic binary BH merger is not trivial given the large space of parameters on which this quantity depends. We have shown that the results of recent simulations combined with basic but exact considerations derived from the EMRL allow us to model this quantity with a simple analytic expression in the case of BH binaries having unequal masses and unequal spins which are aligned with the orbital angular momentum. When compared with all other estimates coming either from numerical calculations or from approximation techniques, the estimates of the 2D fit show differences which are of few percent at most. \\bigskip \\noindent We % thank A. Buonanno, T. Damour, S. A. Hughes, L. Lehner, A. Nagar, and B. S. Sathyaprakash for discussions. We are grateful to D. Merritt for pointing out an error in the interpretation of our results. The computations were performed on the supercomputers at AEI, LITE, LSU, LONI and NCSA. Support comes also through the DFG grant SFB/TR% ~7." }, "0710/0710.1340_arXiv.txt": { "abstract": "Optically thin two-temperature accretion flows may be thermally and viscously stable, but acoustically unstable. Here we propose that the O-mode instability of a cooling-dominated optically thin two-temperature inner disk may explain the 23-day quasi-periodic oscillation (QPO) period observed in the TeV and X-ray light curves of Mkn~501 during its 1997 high state. In our model the relativistic jet electrons Compton upscatter the disk soft X-ray photons to TeV energies, so that the instability-driven X-ray periodicity will lead to a corresponding quasi-periodicity in the TeV light curve and produce correlated variability. We analyse the dependence of the instability-driven quasi-periodicity on the mass (M) of the central black hole, the accretion rate ($\\rm{\\dot{M}}$) and the viscous parameter ($\\alpha$) of the inner disk. We show that in the case of Mkn~501 the first two parameters are constrained by various observational results, so that for the instability occurring within a two-temperature disk where $\\alpha=0.05-1.0$, the quasi-period is expected to lie within the range of 8 to 100 days, as indeed the case. In particular, for the observed 23-day QPO period our model implies a viscosity coefficient $\\alpha \\leq 0.28$, a sub-Eddington accretion rate $\\dot{M} \\simeq 0.02\\,\\dot{M}_{\\rm Edd}$ and a transition radius to the outer standard disk of $r_0 \\sim 60 \\,r_g$, and predicts a period variation $\\delta P/P \\sim 0.23$ due to the motion of the instability region. ", "introduction": "Strong variability is one of the common observational properties of TeV emitting Active Galactic Nuclei (AGNs) \\cite{cat99}. In many cases, the highly variable high-energy { gamma-rays and the X-rays appear to be correlated with no time delays evident on day-scales}, suggesting that the $\\gamma$-rays { may} result from inverse Compton upscattering of lower energy photons. { The first TeV blazar, for example, that was observed simultaneously in multiple bands from radio to TeV gamma-rays is Mrk~421. The first campaign, conducted in 1994 \\cite{mac95} on Mrk~421, revealed correlated variability between the keV X-ray and TeV gamma-ray emission. The gamma-ray flux varied by an order of magnitude on a timescale of 2 days and the X-ray flux was observed to double in 12 hr. On the other hand, the high-energy gamma-ray flux observed by EGRET, as well as the radio and UV fluxes, showed less variability than the keV or TeV bands. Another multiwavelength campaign on Mrk~421 performed in 1995 revealed another coincident keV/TeV flare \\cite{buk96,taka96}. The UV and optical bands also showed correlation during the flares. With the detection of TeV emission from Mrk~501 \\cite{qui96}, several multiband campaigns were organized on Mrk~501 to verify whether such a behavior is a general property of TeV-emitting blazars or whether it is unique to Mrk 421.} The gamma-ray blazar Mkn~501, detected as a strong TeV emitter in 1995 \\cite{qui96}, is one of the closest ($z = 0.0337$) and brightest BL Lacertae objects. Historically (i.e., prior to 1997), its spectral energy distribution (SED) $\\nu\\,F_{\\rm \\nu}$ resembles that of X-ray selected BL Lac objects, having a peak in the extreme UV--soft X-ray energy band \\cite{kat99,sam96}. { Earlier} optical observations of Mkn~501 have shown variations of up to $1.^{m}3$ and polarized emission up to $P_{\\rm opt} \\simeq 7 \\%$ \\cite{fan99}. During its active state in 1997, where Mkn~501 was monitored in the 2-10 keV X-ray band by the all sky monitor(ASM) on board RXTE and in the TeV energy band by several Cherenkov telescope groups \\cite{aha99a,cat97,dja99,kraw00,pro97,rxte99}, both X-rays and TeV gamma-rays increased by more than 1 order of magnitude from quiescent level \\cite{cat97,pia98}. Analysis of the X-ray and TeV data showed, that the variations were strongly correlated between both bands, { yet only marginally correlated with the optical UV band} \\cite{cat97,dja99,kraw00,pet00}. { While the synchrotron emission peaked below 0.1 keV in the quiescent state, in 1997 it peaked at ~100 keV. This is the largest shift ever observed for a blazar \\cite{pia98}. Earlier investigations of the 1997 April flare in Mkn~501 \\cite{pia98}, showed that its origin may be related to a variation of $\\gamma_{max}$ together with an increased luminosity and a flattening of the injected electron distribution.} During the 1997 high state, the X-ray and TeV light curves displayed a quasi-periodic signature \\cite{hay98,kra99,pro97}. A detailed periodicity analysis, based on the TeV measurements obtained with all Cherenkov Telescopes of the HEGRA-Collaboration and the X-ray data with RXTE, was performed by Kranich et al. \\cite{kra99,kra01} and more recently also by Osone \\cite{oso06} including Utah TA data. The results indeed strongly suggest the existence of a 23 day periodicity, with a combined probability of $p \\simeq 3 \\times 10^{-4}$ in both the TeV and the X-ray light curves covering the same observational period \\cite{kra99,kra01}. \\footnote{{ Note that no QPOs have been seen by MAGIC during 1998-2000, when the source was not very active \\cite{alb07}, suggesting that the QPOs only occur during a very active stage.}} { Rieger and Mannheim ~(2000) have shown that the origin of these QPOs may be related to the presence of a binary black hole in the center of Mkn~501. While this may well be possible, we explore here an alternative scenario, where the observed QPOs are related to accretion disk instabilities.} In a seminal paper, Shapiro et al.~(1976) (SLE) have pointed out that a hot (Compton cooling-dominated) optically thin two-temperature accretion disk may be present in the inner region of the standard optically thick disk \\cite{sha73}, whenever the SLE inequality $\\alpha^{1/4} \\dot{M}_{*}^{2}M_{*}^{-7/4} \\geq 0.6$ is satisfied, where $\\dot{M}_{*} = \\dot{M}/(10^{17} \\rm{g} \\rm{s}^{-1})$, $M_{*}=M/(3 M_{\\odot})$, $\\dot{M}$ is the accretion rate of the disk, $M$ the mass of the central black hole, and $\\alpha$ the viscosity parameter, constrained by the model to lie within the range $ 0.05 < \\alpha < 1$. Later work \\cite{pir78,pri76} has shown that the SLE configuration might be thermally unstable (although less than the standard disk) if the simple standard viscosity prescription is employed. However, relatively small changes in the viscosity law can already ensure a stable configuration \\cite{pir78}, in particular, if stabilizing effects of a disk wind are fully taken into account. A kind of SLE two-temperature disk structure may thus well exist in the inner region of BL Lac type objects \\cite{wan91,wanu91,cao03} and provide a possible explanation for the X-ray and TeV variability phenomenon in Mkn~501. For, firstly, Compton processes in a inner two-temperature disk with electron temperature $T_e$ of about $10^{9}$ K (and ion temperature one or two orders of magnitude higher) can produce (steep) X-ray power-law spectra, in contrast to the standard disk model that can only produce emission up to the optical-UV band \\cite{sha76,wan91}. Secondly, analysis of the linear stability of an optically thin two-temperature disk around a compact object shows that the disk is subject to a radial pulsational instability (inertial-acoustic mode instability) \\cite{wu97}. This possible mode of pulsational overstability, in which radial disk oscillations with local Keplerian frequencies become unstable against axisymmetric perturbations, occurs if the viscosity coefficient increases sufficiently upon compression \\cite{kat78}. In this case, thermal energy generation due to viscous dissipation increases during compression, leading to amplification of the oscillations analogous to the role played by nuclear energy generation in stellar pulsations. As we demonstrate below, the occurrence of such a type of disk oscillation may well account for some quasi-periodic time variability in AGNs in a way similar to Galactic black hole candidates \\cite{blu84,hon92,man96,yan97}. ", "conclusions": "The X-ray to TeV spectra of TeV blazars have been often interpreted within a one-zone SSC model, e.g. \\cite{blo96,der97,mas97,tav98,tav01}. However, if a two-temperature disk structure is present in some of these sources, the real situation may be more complex as the X-ray radiation from a two-temperature disk, for example, may represent an additional, non-negligible source of seed photons for the inverse Compton scattering to TeV energies, in particular during active source stages, likely to be associated with changes in accretion history. Hence our model assumes, that a pulsational instability occurring within a two-temperature disk leads to observable, periodic variations of its X-ray radiation field. Part of this periodically modulated X-ray emission will enter the jet and (in addition to direct synchrotron photons) serves as seed photons for Compton-upscattering to TeV energies { similar as in \\cite{der92,der93}}. Our model thus takes it that both, a direct synchrotron self Compton and an external Compton contribution are relevant for modelling the SED of Mkn~501, the seed photons for external Compton-upscattering consisting of both, the infrared-optical seed photons from the (quasi-steady) SS disk component and the variable X-ray photons from the two-temperature disk component. A related, but more simpler scenario, assuming the seed photons for inverse Compton scattering to be provided by a (direct) synchrotron plus an quasi-steady flux component comparable to the observed infrared-optical flux, has been proposed by Pian et al.~(1998) in order to account for the different degrees of SED variations of Mkn~501 at X-ray and sub-X-ray energies during the April 1997 outburst (see also \\cite{ghi98,kat99}). Detailed analysis indeed suggests that one-component SSC models cannot fit both the April 1997 SEDs and the lightcurves from X-ray to TeV, and that (at least) an additional, moderately variable low energy component contributing in the energy range between 3-25 keV is required \\cite{kraw02,mas04}. A similar conclusion seems to hold for the strong X-ray outburst observed in July 1997 \\cite{lam98}. Interestingly, observations in 1998 also provide evidence for an additional component in the optical regime, possible associated with the SS disk component in our hybrid disk model { (see also \\cite{kat01} for an alternative interpretation)}: As shown by Massaro et al.~\\cite{mas04} optical to X-ray data taken in June 1998 indicate that the optical spectrum is steep and does not match the low energy extrapolation of the X-ray spectrum, hence suggesting the presence of different emission components in the optical and in the X-ray regime as naturally expected in our model. Monitors in both the X-rays and the TeV emission show evidence for a 23-day periodicity during the 1997 high state. As demonstrated above the 23-day period in the X-ray light curve may be caused by a pulsational instability in two-temperature accretion disk, and via the inverse Compton process result in the same periodicity in the TeV light-curve. A pulsational instability occurring in a two-temperature disk with transition radius $r_0 \\sim (48-65)\\, r_{g}$ will result in a recurrence timescale of 8 to 100 days. Based on the observed TeV variability we have employed a characteristic black hole mass of $9 \\times 10^7 M_{\\odot}$ in our calculations. We note that quite different central mass estimates for Mkn~501 have been claimed in the literature, ranging from several times $10^7$ (mainly based on high energy emission properties) up to $10^9\\,M_{\\odot}$ (based on host galaxy observations), see e.g. \\cite{cao02,dej99,fal02,fan05,rie03}. However, as shown by Rieger \\& Mannheim~\\cite{rie03} uncertainties associated with host galaxy observations may easily lead to an overestimate of the central black hole mass in Mkn~501 by a factor of three and thus reduce the implied central mass to $\\simeq (2-3)\\times 10^8\\,M_{\\odot}$, { a value in fact recently confirmed by an independent analysis of central mass constraints derived from host galaxy observations \\cite{woo05}.} Moreover, as argued by the same authors some of the apparent disagreement in central mass estimates may possibly be resolved if a binary black hole system exists in the center of Mkn~501, see also \\cite{rie00,rie03,vil99}, similar as in the case of OJ~287 \\cite{sil88}. For example, if Mkn~501 harbours a binary system with a more massive primary black hole of $\\lppr 10^{9} M_{\\odot}$ and a less massive (jet-emitting) secondary black hole of $\\sim 10^{8}\\,M_{\\odot}$, the mass ratio $\\rho=m/M$ would be of order 0.1, which may compare well with the result $\\rho<0.25$ estimated for OJ~287 \\cite{liu02}. While our characteristic black hole mass employed falls well within the above noted range, we note that the SLE pulsational instability model may still work successfully, if a higher black hole mass is used. For example, if one adopts $M = 3 \\times 10^{8} M_{\\odot}$ and $P=23$ days, one obtains $\\alpha \\leq 0.07$, $r_{0*}=46$ and $\\dot{M}\\simeq 0.008 \\dot{M}_{\\rm Edd}$. Our analysis is based on a specific disk model (SLE) which is open to questions, in particular with respect to its possible stability properties. We note however, that a relatively small change in the usually employed viscosity description may already lead to a thermally stable configuration \\cite{pir78}. On the other hand, it may as well be possible that the SLE configuration represents a quasi-transient phenomenon associated with those changes in accretion history that probably initiate the high states. An alternative (inner) disk configuration of interest may be represented by an optically thin two-temperature ADAF solution \\cite{nar98,yi99}. Such a configuration can exist for accretion rates $\\dot{m}= \\dot{M}/\\dot{M}_{\\rm Edd}$ below a critical rate $\\dot{m}_{\\rm crit} \\simeq 0.3\\,\\alpha^2 \\simeq 0.019$, where the canonical ADAF value of $\\alpha =0.25$ has been employed \\cite{nar98}. ADAFs are generally less luminous than a standard disk, with the typical ADAF luminosity given by $L_A \\simeq 0.02\\,(\\dot{m}/\\alpha^2)\\,\\dot{M}\\,c^2$ \\cite{yi99}. Using the constraints above, the possible ADAF luminosity for Mkn~501 becomes $L_A \\leq 1.3 \\times 10^{43}$ erg/s, which is already about an order of magnitude smaller than required by the X-ray analysis. This suggests that -- at least during its high state -- an optically thin ADAF is not a viable option for Mkn~501. Based on Eq. (7) we can estimate the variation rate of the period due to the motion of the instability \\begin{eqnarray} \\delta P/P = {\\frac{3}{2}} P{\\frac{1}{r}} {\\frac{\\partial{r}}{\\partial{t}}}\\,. \\end{eqnarray} Using $v_{r} = {\\frac{\\partial{r}}{\\partial{t}}}$, $\\dot{M} = 4 \\pi \\rho h r v_{r}$ and Eqs. (4) and (5), one finds \\begin{eqnarray} \\delta P/P =0.44 ~\\alpha^{-7/6}~\\dot{M_{24}}^{5/6}~M_{7}^{-5/6}~r_{*}^{-1/4} \\zeta^{-1/6} \\end{eqnarray} For $\\alpha$ = 0.28 the relevant parameters result in $\\delta P/P= 0.23$. From the period analysis performed by Kranich et al.~\\cite{kra99} a $3\\sigma$ deviation in period corresponds to 6.67 days. If we take this deviation as the intrinsic variation on the periodicity (P), then a $\\delta P/P$ of ${\\frac{6.67}{23}}=0.29$ can be estimated from the results by Kranich et al.~\\cite{kra99}. As this estimate assumes that the deviation of the period is only affected by the motion of the instability, while it may in fact be caused by more than one effect, our theoretical $\\delta P/P$ should not be greater than the observational results. We conclude that the observed 23-day QPOs in Mkn~501 might be caused by the instability of a two-temperature accretion disk. The model presented here may thus offers an alternative explanation to the binary-driven helical jet model of Rieger \\& Mannheim~(2000). Comprehensive computational modelling of the pulsational instability in a two-temperature, cooling dominated disk will be essential to verify this in more detail. Our model predicts that a period correlation in the X-ray and $\\gamma$-ray should always be present during an active source stage, while the period of the QPOs may vary as the instability region could change from one high state to the other." }, "0710/0710.4086_arXiv.txt": { "abstract": "{ Based on the results of applying the extended ADC emission model to three Z-track sources: GX\\th 340+0, GX\\th 5-1 and Cyg\\th X-2, we propose an explanation of the Z-track sources in which the Normal and Horizontal Branches are dominated by the increasing radiation pressure of the neutron star. The emitted flux becomes several times super-Eddington at the Hard Apex and Horizontal Branch and we suggest that the inner accretion disk is disrupted by this and that part of the accretion flow is diverted vertically. This position on the Z-track is exactly the position where radio emission is detected showing the presence of jets. We thus propose that high radiation pressure is a necessary condition for the launching of jets. We also show that flaring must consist of unstable nuclear burning and that the mass accretion rate per unit emitting area of the neutron star $\\dot m$ at the onset of flaring agrees well with the critical theoretical value at which burning becomes unstable. ", "introduction": "% \\label{sect:intro} The Z-track sources are the brightest group of Galactic low mass X-ray binaries (LMXB) containing a neutron star persistently emitting at the Eddington luminosity or several times this. The sources trace out a Z-shape in hardness-intensity (Hasinger et al. 1989) clearly showing that strong physical changes take place, probably at the inner disk and neutron star, but a convincing explanation of the Z-track phenomenon does not exist. The majority of LMXB are of the Atoll class which show somewhat different shapes in hardness-intensity which are also not understood, and neither is the relation between the two classes making our understanding of LMXB very incomplete. Moreover, it is well-known that the Z-track sources are detected as radio emitters, but in one branch only, the horizontal branch. Not only is radio detected, but striking results from the VLA show the release of a massive radio condensation from the source Sco\\th X-1 (Fomalont et al. 2001). Because radio is detected essentially in one branch only, the sources offer the possibility of determining the conditions found in this branch distinguishing it from the other two branches, and so finding the conditions necessary for jet formation. Possible ways of understanding the Z-track sources are by theoretical approaches, timing studies or spectral studies. A theoretical model for the Z-track sources was produced by Psaltis et al. (1995) based on a magnetosphere of the neutron star and the changing properties and geometry of this as the mass accretion rate changed. However, the model assumed that the Comptonized emission observed in the spectra (of all LMXB) originated in a small central region close to the neutron star, and this is inconsistent with our more recent measurements of Comptonizing region size (Church \\& Ba\\l uci\\'nska-Church 2004, below). Extensive timing studies have been made to investigate QPO variations around the Z-track (e.g. van der Klis et al. 1987), but this has not revealed the nature of the Z-track. Previous spectral fitting has applied the Eastern model (Done et al. 2002, Agrawal \\& Sreekumar 2003; di Salvo et al. 2002) which assumes the X-ray emission consists of disk blackbody emission plus non-thermal emission from a small central Comptonizing region. However, our work over a period of 10 years with the dipping class of LMXB provides strong evidence that the source of Comptonized emssion, the ADC, is very extended, typically having a radial extent that is 15\\% of the accretion disk size, but increasing with source luminosity, and this is inconsistent with the Eastern model. As a result we have proposed the ``extended ADC'' emission model consisting of blackbody emission from the neutron star plus Comptonized emission from an extended ADC (Church Ba\\l uci\\'nska-Church 1995). Moreover, the pattern of parameter changes obtained by fitting the Eastern model to the Z-track sources is not very easy to interpret and does not immediately suggest a convincing physical explanation. Thus in the present work, we take the approach of applying the extended ADC model for the first time to the Z-track sources and we present the results of applying this model to the sources GX\\th 340+0, GX\\th 5-1 and Cygnus\\th X-2. \\begin{figure*}[!ht] % \\begin{center} \\includegraphics[width=60mm,height=140mm,angle=270]{church_2007_01_fig1a} % \\includegraphics[width=60mm,height=80mm,angle=270]{church_2007_01_fig1b} % \\caption{Top: Background-subtracted and deadtime-corrected PCA light curve of the 1997 September observation of GX\\th 340+0 with 64 s binning. Bottom: the corresponding variation of hardness ratio \\hbox{(7.3 -- 18.1 keV)/(4.1 -- 7.3 keV)} with intensity.} \\label{} \\end{center} \\end{figure*} ", "conclusions": "We show that the radiation pressure of the enitting part of the neutron star is very strong at the hard apex and horizontal branch of the Z-track in three sources, exactly correlating with the parts of the Z-track where radio emission is observed showing the presence of jets, and we suggest that strong radiation pressure is a necessary condition for jet formation." }, "0710/0710.1947_arXiv.txt": { "abstract": "% {} {We study the variability of the Fe 6.4 KeV emission line from the Class I young stellar object Elias 29 in the $\\rho$ Oph cloud.} {We analysed the data from Elias 29 collected by \\xmm\\ during a nine-day, nearly continuous observation of the $\\rho$ Oph star-forming region (the Deep Rho-Oph X-ray Observation, named {{\\sc Droxo}}). The data were subdivided into six homogeneous time intervals, and the six resulting spectra were individually analysed} {We detect significant variability in the equivalent width of the Fe 6.4 keV emission line from Elias 29. The 6.4~keV line is absent during the first time interval of observation and appears at its maximum strength during the second time interval (90 ks after Elias 29 undergoes a strong flare). The X-ray thermal emission is unchanged between the two observation segments, while line variability is present at a 99.9\\% confidence level. Given the significant line variability in the absence of variations in the X-ray ionising continuum and the weakness of the photoionising continuum from the star's thermal X-ray emission, we suggest that the fluorescence may be induced by collisional ionisation from an (unseen) population of non-thermal electrons. We speculate on the possibility that the electrons are accelerated in a reconnection event of a magnetically confined accretion loop, connecting the young star to its circumstellar disk.} {} ", "introduction": "\\label{sec:intro} The X-ray emission from young stellar objects (YSOs) at CCD-resolution is usually modelled as thermal emission from a hot plasma in coronal equilibrium, with higher characteristic temperatures than observed in older and less active stars. An interesting deviation from a pure thermal X-ray spectrum is the presence of fluorescent emission from neutral (or weakly ionised) Fe as shown by the presence of the 6.4 keV line. This was first detected by \\citet{ikt01} in the X-ray emission of the YSO YLW16A in $\\rho$-Oph, during a large flare: in addition to the Fe\\,{\\sc xxv} complex at 6.7 keV, a 6.4 keV emission line was clearly visibe. Such fluorescence line is produced when energetic X-rays photoionise cold material close to the X-ray source, and it is therefore a useful diagnostic tool of the geometry of the X-ray emitting source and its surroundings. Since 2001, detections of the Fe K fluorescent emission line at 6.4-keV in the spectra of YSOs have been reported by a number of authors. \\citet{tfg+05} has identified seven sources with an excess emission at 6.4 keV among 127 observations of YSOs within the COUP observation of Orion; \\citet{fms+05} report 6.4 keV fluorescent emission in Elias 29 in $\\rho$-Oph both during quiescent and flaring emission, unlike all other reported detection of Fe fluorescent emission in YSOs that were made during intense flaring; \\citet{gfm+07} have detected Fe 6.4 keV emission from a low-mass young star in Serpens, during an intense, long-duration flare. Recently, \\cite{sc2007} have reported intense Fe fluorescent emission in the spectrum of V 1486 Ori during a strong flare, when the plasma reached a temperature in excess of 10 keV. The 6.4 keV fluorescent line has been detected in different classes of X-ray emitters: X-ray binaries, active galactic nuclei (AGNs), massive stars, supernova remnants, and the Sun itself during flares. In the case of the Sun, the fluorescing material is the solar photosphere, in the YSOs, however, indications are that the material in the circumstellar disk and its related accretion structures could be responsible for the fluorescence. The typical equivalent width of the 6.4 keV emission line in the studies mentioned above is of the order of 150~eV and is too large to be explained with fluorescent emission in the stellar photosphere or in diffuse circumstellar material (e.g. \\citealp{tfg+05}; \\citealp{fms+05}). This scenario implies that the disk is ``bathed'' in high-energy X-rays emitted by the star, with significant astrophysical implications; for instance, X-rays, in addition to cosmic rays, would play an important role in photoionising the circumstellar material around young star and thus in coupling the gas to the ambient magnetic field (as suggested by e.g. \\citealp{gfm00}). \\citet{cbt+02} suggest that the ``hot'' component they observe in the disk, in the infrared, is heated by the stellar high-energy emission. \\begin{figure*} \\begin{center} \\leavevmode \\epsfig{file=7899fg01.ps, width=17.0cm, bbllx=0,bblly=190,bburx=842,bbury=1000, angle=270, clip=} \\caption{The light curve of Elias 29 over the nine days of the \\droxo\\ observation from PN ({\\em top}), MOS1 ({\\em bottom-left}), and MOS2 ({\\em bottom-right)}. The line with error bars gives the background-subtracted light curve of the source, while the thinner line without error bars gives the total counts (source plus background). The 6 time intervals that we selected for the spectral analysis, on the basis of the PN data, are indicated.} \\label{fig:lc} \\end{center} \\end{figure*} Observations of the time variability of the Fe 6.4 keV emission in YSOs would provide useful constraints on the geometry and sizes of the star-accretion disk system and of other circumstellar structures such as funnel flows, jets, or wind columns. We present here the results of a time-resolved spectral study of the X-ray emission of Elias 29, during $\\sim 9$ days of nearly continuous observation by \\xmm\\ in the context of the ultra-deep observation of $\\rho$-Oph, named \\droxo\\ (from Deep Rho-Oph X-ray observation -- \\citealp{psf+07}). We investigated the presence of variations in its strong Fe 6.4 keV emission over the 9-day time scale covered by the observations. This paper is structured as follows. After a summary, below, of the properties of Elias 29, the observations and data analysis are briefly presented in Sect.\\,\\ref{sec:obs}. Results are summarised in Sect.\\,\\ref{sec:res}; the simulations carried out to assess the reliability of the line detections and the significance of its variability are described in Sect.\\,\\ref{sec:sim}. The results are discussed in Sect.\\,\\ref{sec:disc}. \\subsection{Elias 29 (GY214)} Elias 29 (16:27:09.4, $-$24:37:18.9) with a bolometric luminosity $L = 26-27.5~L_{\\sun}$ (\\citealp{bak+01}; \\citealp{nts06}) is the most luminous Class I YSO in the $\\rho$-Oph cloud. \\cite{mhc98} used the luminosity in the Br$\\gamma$ line to determine the object's accretion luminosity at $L_{\\rm acc}$ = $15-18 L_{\\sun}$, which makes it the source with the highest accretion luminosity in their sample. More recently, \\citet{nts06} used the luminosity of the hydrogen recombination lines to derive an accretion luminosity of 28.8~$L_{\\sun}$. Using millimeter interferometric observations, \\citet{bhc+02} resolved the emission from the disk and the envelope surrounding Elias 29, showing that the disk is in a relatively face-on orientation ($i < 60^{\\circ}$), which explains many of the remarkable observational features of this source, such as its flat spectral energy distribution, its brightness in the near-infrared, the extended components found in speckle interferometry observations, and its high-velocity molecular outflow. Their best-fitting disk model has an inner radius of 0.01 AU, outer radius of 500 AU, and a mass $M = 0.012 M_{\\sun}$. Elias 29 was previously observed in X-rays with ASCA, \\chandra\\, and \\xmm. In the \\chandra\\ observation (\\citealp{ikt01}), the source quiescent phase is characterised by a temperature of 4.3~keV and luminosity of $2.0 \\times 10^{30}$~\\es, fully consistent with the values derived from the subsequent \\xmm\\ observations by \\citet{ogm05}: $kT = (3.6-5.1)$ keV, $N({\\rm H}) = (4.4-5.3) \\times 10^{22}$ cm$^{-2}$, $Z=(0.8-1.3)~Z_{\\sun}$, and $L_{\\rm X} = 2.8 \\times 10^{30}$ \\es. The source was seen flaring during one of the ASCA observations (\\citealp{kkt+97}) and during the \\chandra\\ observation. The two flares had similar intensity and duration with an $e$-folding time of $\\sim 10$~ks (\\citealp{tik+00}; \\citealp{ikt01}). ", "conclusions": "So far, the Fe 6.4 keV emission in YSOs has been explained in terms of fluorescent emission from the photoionised (colder) material in the circumstellar disk. This scenario, however, cannot easily explain the observed variability of the Fe 6.4 keV emission in Elias 29, which occurs in the absence of signficant variations of the observed X-ray continuum. The equivalent width of the line at its maximum strength of $\\sim 250$ eV is also not easily reconciled with a photoionisation scenario. An alternative line-formation mechanism is collisional excitation by a population of non-thermal electrons. We suggest that these electrons could be accelerated by magnetic reconnection events in the accretion tubes that connect the star to its circumstellar disk. The electrons are decelerated in situ by the accreting material, ionising it and causing the observed Fe 6.4 keV emission." }, "0710/0710.0566_arXiv.txt": { "abstract": "We present new observations of the fundamental ro-vibrational CO spectrum of V1647 Ori, the young star whose recent outburst illuminated McNeil's Nebula. Previous spectra, acquired during outburst in 2004 February and July, had shown the CO emission lines to be broad and centrally peaked---similar to the CO spectrum of a typical classical T Tauri star. In this paper, we present CO spectra acquired shortly after the luminosity of the source returned to its pre-outburst level (2006 February) and roughly one year later (2006 December and 2007 February). The spectrum taken in 2006 February revealed blue-shifted CO absorption lines superimposed on the previously observed CO emission lines. The projected velocity, column density, and temperature of this outflowing gas was 30 km s$^{-1}$, $3^{+2}_{-1}\\times10^{18}$ cm$^{-2}$, and 700$^{+300}_{-100}$ K, respectively. The absorption lines were not observed in the 2006 December and 2007 February data, and so their strengths must have decreased in the interim by a factor of 9 or more. We discuss three mechanisms that could give rise to this unusual outflow. ", "introduction": "McNeil's Nebula was recently illuminated by the outburst of V1647 Ori (McNeil 2004), a young star that is embedded in the Lynds 1630 dark cloud and coincides with the 850$\\micron$ continuum source OriBsmm55 (Mitchell et al. 2001). V1647 Ori has a flat SED in the mid-infrared, and is thus a Class I young stellar object (YSO; Andrews et al. 2004). V1647 Ori underwent a similar outburst as recently as 1966 (Aspin et al. 2006), indicating that V1647 Ori is also an EXor. Such pre-main sequence stars undergo eruptive events that dramatically increase their luminosity for periods of months to years (Hartmann 1998). The outbursts are thought to be triggered by a rapid increase in the stellar accretion rate (Hartmann \\& Kenyon 1996). The eruption in November 2003 of V1647 Ori lasted two years. During the outburst, the star brightened by a factor of 50 in X-rays (Kastner et al. 2006), a factor of 250 in the red (6 mag in the $R_C$ band; Fedele et al. 2007a; Brice\\~{n}o et al. 2004), a factor of 15 in the near-IR($\\sim$3 mag in the $J$, $H$, and $K$ bands; Reipurth \\& Aspin 2004), and a factor of $\\sim$15 at wavelengths from $3.6\\micron$ to $70\\micron$ (Muzerolle et al. 2005). From the overall brightening of the source, Muzerolle et al. (2005) concluded that the bolometric luminosity increased by a factor of 15 to 44$L_\\sun$ (see also Andrews et al. 2004) and that the stellar accretion rate increased from $\\sim10^{-7} M_{\\sun}$ yr$^{-1}$ to $\\sim10^{-5} M_{\\sun}$ yr$^{-1}$. Similarly, Gibb et al. (2006) inferred a stellar accretion rate of $3-6\\times10^{-6} M_{\\sun}$ yr$^{-1}$ from the luminosity of the Br$\\gamma$ emission one year later. This is somewhat larger than the typical accretion rate of a young low mass star ($10^{-8}-10^{-7} M_{\\sun}$ yr$^{-1}$; Bouvier et al. 2007), yet lower than is expected for a star of the FUor type ($\\sim10^{-4} M_{\\sun}$ yr$^{-1}$; Hartman \\& Kenyon 1996). During the onset of the outburst of V1647 Ori, observations of atomic lines with P Cygni profiles provided evidence for a hot ($T\\sim$10,000 K), high velocity ($v = -400$ km s$^{-1}$) wind (Brice\\~{n}o et al. 2004; Reipurth et al. 2004; Vacca et al. 2004; Walter et al. 2004; Ojha et al. 2006; Fedele et al. 2007a) with a mass-loss rate of $\\dot{M}_{\\rm wind}$=4$\\times$10$^{-8}$ \\Mdot (Vacca et al. 2004). This mass loss rate is much lower than that of the typical FUor (Hartmann \\& Kenyon 1996) and comparable to that of a classical T Tauri star (cTTS; Hartigan et al. 1995). The absorption component of the P Cygni profile of several lines (e.g. Pa$\\beta$) disappeared within a few months following the peak of the outburst in early 2004 (Gibb et al. 2006). However, P Cygni structure in the H$\\alpha$ profile indicated that a weaker wind continued throughout the outburst phase (Ojha et al. 2006; Fedele et al. 2007a). In contrast to the hydrogen and helium lines, the fundamental near-infrared ro-vibrational emission lines of CO, observed 2004 February 27, were broad, centrally peaked, and compatible in their intensity with an excitation temperature of 2500 K (Rettig et al. 2005). The width of the lines was shown to be consistent with Keplerian orbital motion of the gas within the inner disk surrounding the central star, similar to the broad emission line profiles that are observed around cTTSs and Herbig Ae/Be stars (HAeBes; Najita et al. 2003; Blake \\& Boogert 2004). A later observation, obtained on 2004 July 30, showed that the CO lines remained broad but the temperature of the gas decreased to 1700 K (Gibb et al. 2006). Neither observation showed any indication of CO in an outflow, as a blue shifted absorption component was not detected. We report followup observations of CO from V1647 Ori, which were acquired 2006 February, 2006 December, and 2007 February. By the time of our initial 2006 observation, V1647 Ori had returned to quiescence and the absorption component in the H$\\alpha$ line profile had disappeared (Fedele et al. 2007a and references therein). Presumably the accretion rate had fallen by two orders of magnitude to its pre-outburst accretion rate as the star faded to its pre-outburst brightness (see Muzerolle et al. 2005). As suggested by the work of Najita et al. (2003), only a continued decrease in CO line intensity from the warm gas in the disk was therefore expected. However, the first post-outburst observation revealed the striking metamorphosis of these lines from centrally peaked emission features to emission lines with blue-shifted absorption. Subsequently, by late 2006 and early 2007, the CO emission lines returned to their original centrally peaked structure, indicating that the production of the outflow diminished within one year of the end of the outburst. In this paper we discuss three scenarios that can give rise to such a phenomenon. ", "conclusions": "During quiescence, V1647 Ori is a class I YSO (Andrews et al. 2004). However, its mass-loss rate during outburst was similar to a strongly accreting cTTS (Vacca et al. 2004). While outflows from cTTSs and HAeBes are common, fundamental ro-vibrational CO emission lines with a blue-shifted absorption component have not been observed around any of the more than 300 such sources observed to date (Najita et al. 2000, 2003; Blake \\& Boogert 2004; Rettig et al. 2006; Brittain et al. 2007; J. Brown private communication). Further, Class I YSOs such as GSS 30 IRS 1, HL Tau and RNO 91 do not show ro-vibrational CO emission lines with blue-shifted absorption components (Pontoppidan et al. 2002, Brittain et al. 2005, and Rettig et al. 2006, respectively). In contrast to these systems, the FUor V1057 Cyg has shown blue-shifted overtone ro-vibrational CO absorption lines (Hartmann et al. 2004). While the similarity of the unusual CO outflows is intriguing, there are important differences between FUors such as V1057 Cyg and EXors such as V1647 Ori. First, FUors have greater accretion rates ($\\sim$10$^{-4}$ \\Mdot) and more extreme winds than EXors (Hartmann \\& Kenyon 1996). Secondly, CO is always and {\\it only} detected in absorption toward FUors, and is thought to originate in the accretion disk (Calvet et al. 1993; Calvet et al. 1991). Despite the significant differences between V1057 Cyg and V1647 Ori, both stars have undergone major eruptions that have generated outflows in CO, suggesting that a different mass-loss mechanism may be at work in these systems than in other young stars. Given the unusual nature of the outflowing CO from V1647 Ori, it is of interest to consider the mechanisms that could give rise to this phenomenon. One possibility is that the absorbing CO condensed out of the hot outflowing wind that was observed during the outburst. There were two outflow components noted in the H$\\alpha$ absorption feature: a variable component at 400 km s$^{-1}$ and a steady component at 150 km s$^{-1}$ (Fedele et al. 2007a). If the lower-velocity gas decelerated at a constant rate to $30$ km s$^{-1}$ (the velocity of the outflowing CO), by the time of our second CO observation in 2004 July the wind would have expanded to 10 AU. There was no evidence of blue-shifted CO absorption on this date (Fig. 1). By 2006 February, when the absorption was observed, the moderate velocity wind would have expanded to nearly 40 AU. However, it seems unlikely that the wind could still retain a kinetic temperature of 700K at a distance of 40AU from the star. Furthermore, it is not clear why CO would condense out of this outflow to reveal warm, blue shifted absorption but not out of any of the other outflows with similar or even greater mass-loss rates. Thus we conclude that this scenario is unlikely. A second possibility is that the CO absorption formed in a shell of material that was swept up by the atomic wind. In this case the absorption did not appear until the column density of material was sufficient to produce measurable absorption, and the heating was the result of the interaction of the wind with the shell. When the mass loss decreased at the end of the outburst phase, this heating was eventually shut down. If the CO/H$_2$ ratio in such a shell was similar to that of a dense molecular cloud, 1.5 $\\times$10$^{-4}$, then the column density of gas was 2$\\times$10$^{22}$ cm$^{-2}$. Adopting a normal interstellar extinction--to--gas ratio, i.e., $A_V=5.6\\times 10^{22} N_{\\rm H}$ mag cm$^2$ atom$^{-1}$ (Bohlin et al. 1978), we find that the observed column density of CO corresponded to $\\sim$20 mags of visible extinction. This is a lower limit, as it is possible that selective dissociation of gas in the nebula could drive down the relative abundance of CO. However, the extinction measured on the line of sight toward V1647 Ori appears to have remained relatively unchanged over the entire course of the outburst at $\\sim$11 mags (Vacca et al. 2004; Gibb et al. 2006), of which 6.5 mags is due to the nebula (Fedele et al. 2007b). There is no evidence for an additional 10--20 mags of extinction toward V1647 Ori, and so we conclude that this scenario is also unlikely. A final scenario we consider is that the CO outflow was launched in response to the reorganization of the stellar magnetic field following the sharp drop in the accretion rate. Pre-main sequence stars tend to have kilogauss magnetic fields which mediate stellar accretion and outflows (e.g. Johns-Krull et al. 2000). This stellar magnetic field truncates the accretion disk where the ram pressure from accretion balances the magnetic pressure from the magnetosphere (Camenzind 1990). Consequently, the truncation radius of the disk, $R_T$, is inversely and nonlinearly proportional to the accretion rate, $\\dot{M}$, and given by $ R_T \\propto \\dot{M}^{-2/7}$ (e.g. Bouvier et al. 2007). Thus the two order of magnitude change in the stellar accretion rate experienced by V1647 Ori would have resulted in the truncation radius being shifted by nearly a factor of four. Two years into the outburst, V1647 Ori rapidly faded by a factor of 40 in the R$_C$-band in just 180 days to return to its pre-outburst level (Fedele et al. 2007a). The mid-infrared flux also returned to its pre-outburst level, indicating that the drop in the optical/near-infrared lightcurve was due to the intrinsic fading of the source (Fedele et al. 2007a). This sharp drop in the luminosity of the source indicates that the accretion rate fell rapidly as the star returned to quiescence. In response to the drop in ram-pressure from the accretion flow, the truncation radius was pushed back by the magnetosphere. Evidence from simulations suggests that disk-magnetosphere systems tend to form outflows when the system is undergoing the greatest amount of dynamical rearrangement (e.g., Fig. 7 in Balsara 2004). It is possible that the realignment of the magnetic field as it pushed out against the circumstellar disk resulted in a warm, shortlived outflow, an outflow that is not observed toward any other cTTSs or HAeBes. While virtually all accreting low-mass stars drive outflows, the CO outflow from V1647 Ori is highly unusual. Indeed, the transformation of centrally peaked ro-vibrational CO emission lines to CO emission lines with blue-shifted absorption is unique. We suggest that the mechanism responsible for producing this outflow is distinct from the one that drives the outflow from typical cTTSs. The coincidence between the rapid fading of V1647 Ori and the subsequent observation of the CO outflow hints at a connection. Better sampling of the fundamental ro-vibrational CO spectrum of EXors as they brighten and fade is crucial for determining whether this coincidence is significant. The rapid and dramatic change in the accretion rate that characterizes the EXor phenomenon provides an important opportunity to study the interplay between stellar accretion, the inner disk, and outflows. This insight is key to reaching a satisfactory theoretical understanding of these events." }, "0710/0710.2563_arXiv.txt": { "abstract": "In this study we compile for the first time comprehensive data sets of solar and stellar flare parameters, including flare peak temperatures $T_p$, flare peak volume emission measures $EM_p$, and flare durations $\\tau_f$ from both solar and stellar data, as well as flare length scales $L$ from solar data. Key results are that both the solar and stellar data are consistent with a common scaling law of $EM_p \\propto T_p^{4.7}$, but the stellar flares exhibit $\\approx 250$ times higher emission measures (at the same flare peak temperature). For solar flares we observe also systematic trends for the flare length scale $L(T_p) \\propto T_p^{0.9}$ and the flare duration $\\tau_F(T_p) \\propto T_p^{0.9}$ as a function of the flare peak temperature. Using the theoretical RTV scaling law and the fractal volume scaling observed for solar flares, i.e., $V(L) \\propto L^{2.4}$, we predict a scaling law of $EM_p \\propto T_p^{4.3}$, which is consistent with observations, and a scaling law for electron densities in flare loops, $n_p \\propto T_p^2/L \\propto T_p^{1.1}$. The predicted ranges of electron densities are $n_p \\approx 10^{9-10}$ cm$^{-3}$ for solar nanoflares at $T_p=1$ MK, $n_p \\approx 10^{10-11}$ cm$^{-3}$ for typical solar flares at $T_p=10$ MK, and $n_p \\approx 10^{11-12}$ cm$^{-3}$ for large stellar flares at $T_p=100$ MK. The RTV-predicted electron densities were also found to be consistent with densities inferred from total emission measures, $n_p=\\sqrt{EM_p/q_V V}$, using volume filling factors of $q_V=0.03-0.08$ constrained by fractal dimensions measured in solar flares. Solar and stellar flares are expected to have similar electron densities for equal flare peak temperatures $T_p$, but the higher emission measures of detected stellar flares most likely represents a selection bias of larger flare volumes and higher volume filling factors, due to low detector sensitivity at higher temperatures. Our results affect also the determination of radiative and conductive cooling times, thermal energies, and frequency distributions of solar and stellar flare energies. ", "introduction": "Scaling laws provide important diagnostics and predictions for specific physical models of nonlinear processes such as self-organized criticality, turbulence, diffusion, plasma heating, and particle accleration. These models have been widely applied in plasma physics, astrophysics, geophysics, and the biological sciences. Here we investigate scaling laws of physical parameters in solar and stellar flares, which should allow us to decide whether solar and stellar flare data are consistent with the same physical flare process. The scaling of solar and stellar flare data has been pioneered by Stern (1992), Feldman et al.~(1995b), and Shibata \\& Yokoyama (1999; 2002), who showed evidence for a nonlinear scaling between the flare volume emission measure $EM_p$ and the flare peak temperature $T_p$. These parameters have been measured in solar flares with instruments like {\\sl Skylab}, {\\sl GOES}, {\\sl Yohkoh/SXT} (Soft X-ray Telescope), and {\\sl RHESSI (Ramaty High Energy Solar Spectroscopic Imager)}, and in stellar flares with {\\sl ASCA}, {\\sl BeppoSAX}, {\\sl Einstein}, {\\sl EUVE}, {\\sl EXOSAT}, {\\sl Ginga}, {\\sl HEAO}, {\\sl ROSAT}, {\\sl Chandra}, and {\\sl XMM-Newton}. Compilations of solar flare parameters have been presented in Aschwanden (1999), while stellar flare parameters were compiled in a recent review by G\\\"udel (2004). In this paper we present for the first time this host of mostly new measurements ``on the same page'' and investigate commonalities and differences between the scaling of solar and stellar flares. In Section 2 we present the statistical correlations found in stellar flare data, while the corresponding counterparts of solar flare data are shown in Section 3. In Section 4 we present theoretical modeling of the data, using the well-known RTV law, the generalization with gravitational stratification and spatially non-uniform heating, the fractal flare volume scaling, and volume filling factor. In Section 5 we discuss the differences between solar and stellar scaling laws, the consistency between two different electron density measurement methods, and a previously derived ``universal scaling law'' for solar and stellar flares. Section 6 summarizes our conclusions. ", "conclusions": "We compiled directly observed parameters from solar and stellar flares, such as the volume peak emission measure $EM_p$, flare peak temperature $T_p$, flare duration $\\tau_f$, and flare length scale $L$ (the latter only for solar flares). A prominent statistical correlation is found between the volume emission measure $EM_p$ and flare peak temperature $T_p$, which scales as $EM_p \\approx T_p^{4.7}$ for both solar and stellar flares. Another recent study demonstrated that the flare volume has a fractal scaling, $V(L)\\propto L^{2.4}$, rather than the generally used Euclidian scaling of $V(L)\\propto L^3$. Applying the RTV scaling law, combined with the fractal volume scaling and the statistical $L-T_p$ correlation $L(T_p) \\propto T_p^{0.9}$, leads directly to a theoretically predicted scaling law of $EM_p \\propto T_p^{4.3}$, which explains the observed correlations in both solar and stellar flares. A second result we find is an unexplained offset by a factor of about 250 between solar and stellar flares at the same temperature, which is likely due to a selection bias for stellar flare events with larger volume filling factors and larger spatial scales. Interestingly, however, this selection bias does not affect the overall $EM-T$ relationship and the lower threshold has a similar functional dependence of $EM_{min} \\propto T_p^{4.7}$, probably because the detector sensitivities are dropping off with a similar function with higher temperatures. A third result is that our model of fractal flare volume scaling provides realistic estimates of volume filling factors, and thus of flare densities. We find that the electron densities in solar flare loops can be predicted based on our fractal scaling in close agreement to the predictions of the RTV law. The agreement of the predicted electron densities with both methods agrees always better than an order of magnitude (Fig.~7), although the absolute magnitude varies by three orders of magnitude between the smallest nanoflares and the largest solar flares, i.e., $n_e \\approx 10^9-10^{12}$ cm$^{-3}$. Since the RTV scaling law (combined with the observed $L \\propto T_p$ correlation) predicts about a linear relationship between the electron densities and flare peak temperatures, i.e., $n_p \\propto T_p$, we expect up to an order of magnitude higher electron densities in the largest stellar flares due to the higher temperature than in solar flares. The determination of correct scaling laws allows us also to infer realistic estimates of the (conductive and radiative) flare cooling times, which can be tested from the e-folding decay time of individual peaks in (solar and stellar) flare light curves. The scaling laws allow us also to eliminate temperature biases in the statistics of the total thermal energy of flares, i.e., $E_T \\propto n_p T_p V \\propto EM_p T_p/n_p$. Since we find that the electron density (corrected for a fractal filling factor) scales approximately as $n_p \\propto T_p^{1.1}$, the thermal energy scales approximately as $E_T \\propto EM_p$, and thus the observed total emission measure $EM_p$ can be used as a good proxy for the thermal flare energy $E_T$. Such unbiased frequency distributions of flare energies $N(E_T)$ permit us then to determine whether there is more energy in large or small flares, an important test for nanoflare heating theories." }, "0710/0710.0799_arXiv.txt": { "abstract": "{ Non thermal emission from galaxy clusters demonstrates the existence of relativistic particles and magnetic fields in the Intra Cluster Medium (ICM). Present instruments do not allow to firmly establish the energy associated to these components. In a few years gamma ray observations will put important constraints on the energy content of non thermal hadrons in clusters, while the combination of radio and hard X-ray data will be crucial to measure the energy content in the form of relativistic electrons and magnetic field. SIMBOL-X is expected to drive an important breakthrough in the field also because it is expected to operate in combination with the forthcoming low frequency radio telescopes (LOFAR, LWA). In this contribution we report first estimates of {\\it statistical properties} of the hard X--ray emission in the framework of the {\\it re-acceleration model}. This model allows to reproduce present radio data for Radio Halos and to derive expectations for future low frequency radio observations, and thus our calculations provide hints for observational strategies for future radio and hard--X-ray combined observations. ", "introduction": "Clusters of galaxies represent the largest virialized structures in the present Universe. Rich clusters have typical total masses of $10^{15} M_{\\odot}$, mostly in the form of dark matter, while $\\sim 5\\%$ of the mass is in the form of a hot ($T \\sim 10^8 K$), tenuous ($n_{gas} \\sim 10^{-3}-10^{-4} cm^{-3}$), X-ray emitting gas. In terms of energy density, the gas is typically heated to roughly the virial temperature, but there is also room to accomodate a non-negligible amount of non-thermal energy. Clusters are ideal astrophysical environments for particle acceleration and cosmic rays (CR) accelerated within the cluster volume are expected to be confined for cosmological times (e.g., Blasi, Gabici, Brunetti 2007, BGB07, for a review). The bulk of the energy of these CRs is expected in protons since they have radiative and collisional life--times much longer than those of the electrons. While present gamma ray observations can only provide upper limits to the average energy density of CR protons in the ICM (e.g. Reimer et al. 2004), evidence of a non-thermal component is in fact obtained from radio observations of a fraction of galaxy clusters showing synchrotron emission on Mpc scales : Radio Halos, fairly symmetric sources at the cluster center, and Radio Relics, elongated sources at the cluster periphery (e.g., Feretti 2005). Although the bulk of present data comes from radio observations, theoretically a substantial fraction of the non thermal radiation is expected from inverse Compton (IC) scattering of the photons of the cosmic microwave background (e.g., Sarazin 1999). Measuring IC emission from clusters in the hard X--rays is extremely important to derive the energy density of emitting electrons and the strength of the magnetic field when these measures are combined with radio data. Despite the poor sensitivity of present and past hard X--ray telescopes, several groups have claimed detection of hard X--ray emission (HXR) in a few massive clusters (e.g., Fusco-Femiano et al.~2004; Petrosian et al.~2006; Rephaeli et al.~2006; see also Rossetti \\& Molendi 2004 and Fusco-Femiano et al.~2007 for a discussion on the strength of the HXR detection in the Coma cluster). Thanks to its sensitivity and capability to perform hard X--ray imaging SIMBOL--X will open a new era in the study of non thermal radiation from galaxy clusters. In this contribution we report first expectations on the Luminosity Functions (LFs) and number counts of HXR from clusters. We calculate only the contribution to the IC spectrum from electrons re-accelerated by turbulence in the ICM which are the responsible for the origin of Radio Halos in the context of the {\\it re-acceleration scenario}. ", "conclusions": "In this contribution we report first expectations for HXRs from galaxy clusters, a more detailed study will be reported in a forthcoming paper. Calculations are performed in the framework of the {\\it re-acceleration model} assuming physical parameters which allow the {\\it re-acceleration model} to reproduce present data of the statistical behaviour of giant Radio Halos. The strength of the magnetic field in the ICM is a crucial parameter in our calculations and we have shown that SIMBOL-X will provide unique constraints. By assuming a value of the magnetic field averaged in Mpc$^3$ volume of $\\approx$0.2$\\mu$G we find that SIMBOL-X will discover HXRs in $\\approx$30--100 clusters at z$\\leq$0.2 ." }, "0710/0710.0367_arXiv.txt": { "abstract": "{ Combined X-ray synchrotron and inverse-Compton $\\gamma$-ray observations of pulsar wind nebulae (PWN) may help to elucidate the processes of acceleration and energy loss in these systems. In particular, such observations provide constraints on the particle injection history and the magnetic field strength in these objects. The newly discovered TeV $\\gamma$-ray source HESS\\,J1718$-$385 has been proposed as the likely PWN of the high spin-down luminosity pulsar PSR\\,J1718$-$3825. The absence of previous sensitive X-ray measurements of this pulsar, and the unusual energy spectrum of the TeV source, motivated observations of this region with \\emph{XMM-Newton}. The data obtained reveal a hard spectrum X-ray source at the position of PSR\\,1718$-$3825 and evidence for diffuse emission in the vicinity of the pulsar. We derive limits on the keV emission from the centroid of HESS\\,J1718$-$385 and discuss the implications of these findings for the PWN nature of this object. } ", "introduction": "Young pulsars drive relativistic winds into their environments, confinement of which leads to the production of extremely broadband emission via the synchrotron and inverse-Compton (IC) processes \\citep[see][for a recent review]{PWN:review}. The most prominent PWN, the Crab Nebula, is detected in all wavebands from the radio to TeV $\\gamma$-rays~\\cite{Whipple:crab}, with the transition from synchrotron to IC emission at $\\sim$1~GeV. The recent increase in sensitivity of ground-based TeV $\\gamma$-ray instruments has led to a rapid increase in the number of putative PWN in this waveband. These objects are characterised by diffuse, typically offset, nebulae around high spin-down luminosity pulsars. The archetype of this new object class is the PWN G\\,18.0---0.7/HESS\\,J1825$-$137. G\\,18.0---0.7 is a $\\sim$5$'$ long asymmetric X-ray synchrotron nebula associated with the middle-aged (characteristic spin-down age $\\tau$=21~kyr) pulsar PSR\\,B1823$-$13~\\cite{XMM:1825}. The IC nebula HESS\\,J1825$-$137 is much larger ($\\sim$100$'$ at 1~TeV) but exhibits energy-dependent morphology, shrinking towards the pulsar at high energies~\\cite{HESS:1825p2,HESS:1825icrc}, suggestive of cooling of the highest-energy (X-ray synchrotron emitting) electrons away from the pulsar. The TeV $\\gamma$-ray source HESS\\,J1718$-$385 was discovered in deep observations of the supernova remnant RX\\,J1713.7$-$3946 using H.E.S.S. in 2004-2005~\\cite{HESS:twopwn}. The absence of other potential counterparts and the relatively compact nature of the source ($9'\\times4'$ rms) make an association with PSR\\,J1718$-$3825 ($8'$ from the centroid of the TeV source) plausible. The TeV source is unusual in its sharply peaked spectral energy distribution (SED), which is similar to that of the $\\gamma$-ray nebula of the Vela pulsar~\\cite{HESS:velax}. The $\\gamma$-ray emission from these objects is commonly attributed to IC scattering of relativistic electrons \\citep[see][for an alternative view]{Horns:VelaX}. In this scenario the spectral break seen at $\\sim$10~TeV in these objects can be interpreted as a signature of electron cooling. However, PSR\\,J1718$-$3825 (estimated distance 4.2~pc) has a characteristic spin-down age (90~kyr) almost an order of magnitude greater than that of the Vela pulsar, making such a high energy break very surprising. The search for a possible X-ray counterpart to HESS\\,J1718$-$385 is important for two reasons: firstly, to verify the identification of the TeV source as the PWN of PSR\\,J1718$-$3825 and secondly, to explore the physical conditions and electron energy distribution in the putative nebula. As no sensitive X-ray observations of the PSR\\,J1718$-$3825/HESS\\,J1718$-$385 region existed, \\emph{XMM-Newton} was used to observe this region in September 2006. ", "conclusions": "The discovery of hard spectrum X-ray emission from the vicinity of PSR\\,J1718$-$3825, and the evidence for a diffuse halo around the pulsar strongly suggest the existence of a synchrotron nebula around this pulsar. This discovery strengthens the association of the $\\gamma$-ray source HESS\\,J1718$-$385 to PSR\\,J1718$-$3825, but the relationship of the X-ray emission to the $\\gamma$-ray source is not straightforward. The overall asymmetry of the nebula with respect to the pulsar is consistent with the idea of SNR expansion into a non-uniform molecular environment \\citep[see for example][]{blondin01:PWN}. The very different morphologies in the two wavebands suggest that either electrons of rather different energies are responsible for the two sources and/or that the magnetic field strength within the nebula is highly non-uniform. As the target for IC emission is the CMBR and other large-scale radiation fields, the IC flux $F_{\\rm IC}$ is simply proportional to the number of radiating electrons, $n_{e}$, whereas the synchrotron flux goes as: $F_{\\rm synch}\\propto B^{2}n_{e}$. In either case the situation may be rather similar to that of HESS\\,J1825$-$137 or indeed HESS\\,J1813$-$178 \\cite{Funk:1813}, with the lifetime of TeV $\\gamma$-ray emitting electrons being longer than the age of the pulsar, and having time to propagate over distances of several parsecs. The SED of the source is presented in Figure~\\ref{fig:sed}. Three features are of note: \\begin{figure}[t] \\centering \\resizebox{0.9\\hsize}{!}{\\includegraphics{Figure3}} \\vspace{-1mm} \\caption{Spectral energy distribution for the pulsar wind nebula of PSR\\,J1718$-$3825. The de-absorbed spectrum of emission from within $1'$ of the pulsar is shown together with limits for diffuse emission from the region covered by HESS\\,J1718$-$385. Three sets of illustrative synchrotron and inverse Compton model curves are shown, based on assumptions of: {\\bf A)} Mono-energetic 70~TeV electrons injected over a $10^{4}$ year period, $B = 5\\mu$G, {\\bf B)} 70 TeV electrons, $B=20\\,\\mu$G, $t=8$ years and {\\bf C)} An electron energy distribution following a power-law ($\\alpha=1.8$) with an abrupt cut-off at 100~TeV. These curves are calculated as described in \\citet{GC:Hinton07}. } \\label{fig:sed} \\end{figure} 1) a hard IC spectrum at TeV energies, with a peak at $\\sim$10~TeV. This suggests that the electrons responsible for this emission are uncooled. This is rather surprising given the spin-down age of the pulsar (90 kyr): cooling on the CMBR alone would result in a spectral break at 2~TeV after 90~kyr, inconsistent with the $\\gamma$-ray data. Indeed ages $>40$ kyrs appear to be excluded by the data. A true age of $\\sim$10~kyr could be explained by a birth period for the pulsar very close to its current period of 75~ms, or breaking deviating significantly from the pure magnetic dipole case (as it seems may commonly be the case, see \\citet{Kramer:Spindown}) \\footnote{We further note that the projected length of the $\\gamma$-ray nebula ($\\sim$10~pc) is roughly half that of HESS\\,J1825$-$137, despite the fact that PSR\\,J1718$-$3825 is apparently a factor 5 older. Another possibility is that HESS\\,J1825$-$137 represents only the youngest part of a larger, softer spectrum, $\\gamma$-ray nebula.}. The shape of the TeV spectrum appears to be consistent with a constant injection of (mono-energetic) $\\sim$70~TeV electrons (curve {\\bf A}), or with a hard power-law (index $\\approx$1.8) with a sharp cut-off around 100 TeV ({\\bf C}). 2) hard spectrum X-ray emission from the pulsar vicinity with a much lower energy flux. The spectrum from within $1'$ of PSR\\,J1718$-$3825 shown in Figure~\\ref{fig:sed} represents the combined flux of pulsar itself and the inner PWN. The pulsar contribution is certainly less than half of the total emission (as is clear from the flux in the annulus surrounding the pulsar) and for typical systems of this type represents only $\\sim$20\\% of the PWN emission in the $>2$keV range \\citep{Kargaltsev:ChandraPWN}. Assuming the PWN is dominant, the hard spectrum suggests either higher electron energies or larger magnetic fields in this region ({\\bf B}) in comparison to those found in the $\\gamma$-ray nebula. The lower flux can be explained if only recently injected electrons are confined in the region around the pulsar. Whilst the two-zone scenario illustrated by curves {\\bf A} and {\\bf B} is clearly grossly oversimplified, is does appear that the data are consistent with the idea that electrons with a relatively narrow energy distribution rapidly escape from a high $B$-field region close to the pulsar into the extended nebula seen in $\\gamma$-rays. Another plausible scenario is that the injection spectrum of electrons has changed significantly over the lifetime of the pulsar. 3) the energy flux level of diffuse X-ray emission from the HESS\\,J1718$-$385 region exceeds the TeV flux by not more than a factor $\\sim$2. The energy distribution of electrons is essentially fixed by the TeV data. The diffuse X-ray limits can therefore be used constrain the $B$-field in the extended nebula to be not much greater than $5\\,\\mu$G (curves {\\bf C}+{\\bf A}), close to the mean value of the ISM. We note that this constraint comes principally from the diffuse limit at 2--4.5 keV which is relatively independent of the assumed absorbing column. In conclusion, the diffuse X-ray emission around XMMU\\,171813.8$-$382517 and HESS\\,J1718$-$385 appear to represent different zones in the PWN of the middle-aged pulsar PSR\\,J1718$-$3825. Future studies of this complex system are certainly well motivated. For example, with the superior angular resolution of Chandra, the contribution of the pulsar itself to the non-thermal emission could be separated from that of the PWN." }, "0710/0710.5774_arXiv.txt": { "abstract": "We review the properties of carbon-sequence ([WC]) Wolf-Rayet central stars of planetary nebulae (CSPNe). Differences between the subtype distribution of [WC] stars and their massive WC cousins are discussed. We conclude that [WO]-type differ from early-type [WC] stars as a result of weaker stellar winds due to high surface gravities, and that late- and early-type [WC] and [WO] stars generally span a similar range in abundances, X(He)$\\sim$X(C)$\\gg$X(O), consistent with a late thermal pulse, and likely progenitors to PG1159 stars. ", "introduction": "% This review discusses the properties of the small fraction of central stars of Planetary Nebulae (CSPNe) which share a spectroscopic appearance with massive, carbon-sequence (WC-type) Wolf-Rayet stars. Massive WC stars are the chemically evolved descendents of initially very massive O stars ($M_{\\rm init} \\geq 25 M_{\\odot}$) exhibiting the C and O products of core helium burning, plus a unique emission line spectral appearance due to fast, dense stellar wind outflows. Such stars are young, with ages of only a few Myr of which several hundred cases are known within the Milky Way, supplemented by thousands more known in external star-forming galaxies (Crowther 2007; Hamann these proc.). CSPNe possessing a similar spectral morphology to WC stars are denoted [WC] and are at a post-Asymptotic Giant Branch (AGB) phase in the late stages of evolution of low or intermediate mass stars ($M_{\\rm init} \\sim 1-5 M_{\\odot}$?) with only a few dozen examples known in the Milky Way plus a handful in the Magellanic Clouds. With respect to normal H-rich CSPNe, the unusual surface chemical composition of [WC] stars apparently results from a late thermal pulse (LTP), causing a H-deficient surface, and likely connection with other H-deficient stars, most notably PG1159 stars (Werner et al. these proc.) ", "conclusions": "We discuss physical and wind properties of Wolf-Rayet CSPNe drawn from the recent literature plus new analyses for a range of subtypes. In general the higher ionization lines seen at earlier spectral type indicates an increased stellar temperature, although very early [WO] subtypes are favoured for hot CSPNe with weak winds, with [WC4--5] subtypes resulting for hot CSPNe with stronger winds. [WC8--9] subtypes correspond to lower temperature CSPNe with strong winds, with much lower temperatures indicated in [WC10] stars. We explain the differences in subtype distribution between Galactic disk CSPNe and massive WC stars as a result of decreased wind densities in the former, owing to increased surface gravities. Massive WC stars in the LMC differ from Galactic WC stars through reduced wind densities caused by lower metallicities, rather than increased surface gravities. There does {\\it not} appear to be a systematic difference between carbon-to-helium mass fractions in late to early-type CSPNe, at least for subtypes for which we are able to employ common diagnostics ([WC9] to [WO1]). Consequently, evolution from late-type [WC] through early-type [WC] and [WO] to PG1159 stars appears to be consistent with most abundance patterns, and expectations for a late thermal pulse. Indeed, we note that similar analysis tools have been applied to the post He-flash system V605 Aql, which has now evolved through to a early-type [WC] spectral type over the past 80 years and shares a similar abundance pattern with X(He):X(C):X(O) = 54:40:5\\% (Clayton et al. 2006)." }, "0710/0710.5542_arXiv.txt": { "abstract": "There are periodic solutions to the equal-mass three-body (and $N$-body) problem in Newtonian gravity. The figure-eight solution is one of them. In this paper, we discuss its solution in the first and second post-Newtonian approximations to General Relativity. To do so we derive the canonical equations of motion in the ADM gauge from the three-body Hamiltonian. We then integrate those equations numerically, showing that quantities such as the energy, linear and angular momenta are conserved down to numerical error. We also study the scaling of the initial parameters with the physical size of the triple system. In this way we can assess when general relativistic results are important and we determine that this occur for distances of the order of $100M$, with $M$ the total mass of the system. For distances much closer than those, presumably the system would completely collapse due to gravitational radiation. This sets up a natural cut-off to Newtonian $N$-body simulations. The method can also be used to dynamically provide initial parameters for subsequent full nonlinear numerical simulations. ", "introduction": "The closest star to the solar system, Alpha Centauri, is a triple system, so is Polaris and HD 188753. Triple stars and black holes are common in globular clusters~\\cite{Gultekin:2003xd, Miller:2002pg}, and galactic disks. Triple black hole mergers can be formed in galaxy merger~\\cite{Valtonen96} and a triple quasar, representing a triple supermassive black hole system has been recently discovered~\\cite{Djorgovski:2007ka}. Full numerical simulations of black holes made possible only in the last couple of years have already produced numerous astrophysically interesting results, among them, the orbital hangup and respect of the cosmic censorship hypothesis for spinning black holes~\\cite{Campanelli:2006uy,Campanelli:2006fg,Campanelli:2006fy}, precession and spin-flips~\\cite{Campanelli:2006fy}, and the discovery~\\cite{Campanelli:2007ew} of large recoil velocities in highly-spinning black hole mergers up to $4,000$ km/s~\\cite{Campanelli:2007cga}. The 2005 breakthroughs in Numerical Relativity~\\cite{Pretorius:2005gq,Campanelli:2005dd,Baker:2005vv}, not only provided a solution to the long standing two-body problem in General Relativity, but it also proved applicable to the black hole - neutron star binaries~\\cite{Faber:2007dv} and recently to the three (and $N$) - black holes systems~\\cite{Campanelli:2007ea}. In general, the solution of three-body problem in Newtonian gravity can be chaotic. There are however, periodic orbits in the problem of three equal masses on a plane. One of the most surprising solution is a figure-eight % orbit. The three bodies chase each other forever around a fixed eight-shaped curve. This was found first by Moore~\\cite{Moore:1993} and discussed with the proof of the existence in Ref.~\\cite{Chenciner:2000}. Heggie~\\cite{Heggie:2000} also estimates the probability for such systems to occur in a galaxy. Because of effects such as the perihelion shift, it was unclear if the figure-eight orbits would exist in a low post-Newtonian expansion, even if it consist of only conservative terms. Imai, Chiba and Asada succeeded in obtaining the figure-eight solution in a first post-Newtonian order approximation by finding the general relativistic corrections to the Newtonian initial conditions. In Ref.~\\cite{Chiba:2006ad} they also estimated the periodic gravitational waves from this system. In Ref.~\\cite{Imai:2007gn} was used the Euler-Lagrange equations of motion in an approximation to first post-Newtonian order. In our paper we instead assume the Hamiltonian formulation to derive the equations of motion. We start from the Hamiltonian given in Ref.~\\cite{Schaefer87} (with typos corrected in our Appendix). We derive the equations of motion in this formalism, which are different from those used in Ref.~\\cite{Imai:2007gn} and have the virtue of explicitly satisfying the constants of motion of the problem, and thus being more amenable to numerical integration. The paper is organized as follows. In Section~\\ref{sec:EOM}, we summarize the equations of motion to be solved numerically in order to obtain the figure-eight orbits. The starting point is the three-body Hamiltonian in the first post-Newtonian approximation. In Section~\\ref{sec:1PN}, we discuss the initial conditions for the figure-eight solutions. We study the scaling relation between the orbital radius and the linear momenta. From this analysis, we can estimate when general relativistic effects are important. In Section~\\ref{sec:2PN}, we extend our calculation to the second post-Newtonian order and in Section~\\ref{sec:DIS}, we summarize the results of this paper and discuss some remaining problems. The 2PN three-body Hamiltonian is explicitly given in the Appendix. Throughout this paper, we use units in which $c=G=1$. ", "conclusions": "\\label{sec:DIS} In this paper we have used the figure-eight orbits as a theoretical lab to test the properties of the low post-Newtonian expansions of General Relativity. We have found that those closed orbits exists for three (and presumably $N$) bodies. We have provided an improved first-post-Newtonian order formalism for deriving the equations of motion that satisfy the Hamiltonian (the linear and angular momenta) constraint to round-off error. The subsequent numerical evolution is well behaved during for more than $t\\sim10,000m$. We have also extended this analysis to the $2PN$ corrections, still giving a conservative system of equations. In the process of finding the figure-eight solutions by trial of different initial momenta we also showed (numerically) the stability of the orbit against small perturbations. This method is particularly useful to determine, dynamically (as an alternative to determine them through families of initial data~\\cite{Campanelli:2005kr}), initial orbital parameters for subsequent full numerical evolution~\\cite{Campanelli:2007ea}, when the holes are close enough that general relativistic effects can no longer be ignored. Note that our method fully takes into account the three-body post-Newtonian interactions unlike other simulations that approximate the problem in successive two-body problems~\\cite{Aarseth:2007wv}. It is interesting to note here that the scaling fits~(\\ref{eq:2PNfit}) give a practical way to determine when relativistic or Newtonian approaches are appropriate. For $\\lambda=1$ we have that the ratio of the first coefficient, $0.01617654493$ (Newtonian) to the second coefficient $0.002017242451$ first-post-Newtonian is nearly $0.12/\\lambda$ and the second coefficient to the third one $0.0002463605227$ (dominated by second-post-Newtonian) is also approximately $0.12/\\lambda$. This indicates that post-Newtonian corrections are important. For $\\lambda=1$ the distance between the initial bodies is $200m$, what indicates that for nearly $67M$ with $M\\approx3m$ the total mass of the system has strong post-Newtonian effects. For $\\lambda \\gg 1$ Newtonian gravity should describe the system accurately, while for $\\lambda<1$ general relativistic effects should be very important, eventually leading to the total collapse of the system. It is interesting to remark here that most of the $N$-body codes use some sort of regularization of the Newtonian gravity for very close encounters \\cite{aarseth-03}, instead the natural way to regularize these close encounters \\cite{Campanelli:2007ea} is given by the General Theory of Relativity, and as we show here, the post-Newtonian corrections are already non-negligible at separations of the order of $100M$. In any case, for most of the astrophysical encounters this is way too short distance, but it can obviously be reached in systems involving black holes and neutron stars." }, "0710/0710.5632_arXiv.txt": { "abstract": "{Homogeneous anisotropic turbulence simulations are used to determine off-diagonal components of the Reynolds stress tensor and its parameterization in terms of turbulent viscosity and $\\Lambda$-effect. The turbulence is forced in an anisotropic fashion by enhancing the strength of the forcing in the vertical direction. The Coriolis force is included with a rotation axis inclined relative to the vertical direction. The system studied here is significantly simpler than that of turbulent stratified convection which has often been used to study Reynolds stresses. Certain puzzling features of the results for convection, such as sign changes or highly concentrated latitude distributions, are not present in the simpler system considered here.} ", "introduction": "The Reynolds stress, described by the correlation of fluctuating velocity components, $Q_{ij} = \\overline{u_i u_j}$, is one of the most important generators of differential rotation in stars (R\\\"udiger \\cite{Ruediger1989}). These stresses have been studied with the help of 3D convection simulations (e.g.\\ Pulkkinen et al.\\ \\cite{Pulkkinenea1993}; Chan \\cite{Chan2001}; K\\\"apyl\\\"a et al.\\ \\cite{Kaepylaeea2004}; R\\\"udiger et al.\\ \\cite{Ruedigerea2005}). These results have revealed some surprising features such as the peaking of the horizontal stress $Q_{xy}$ very close to the equator, and a positive (outward) flux for rapid rotation. Both of these results are at odds with theoretical considerations (Kitchatinov \\& R\\\"udiger \\cite{KitcRued1993}). Furthermore, disentangling of the diffusive (turbulent viscosity) and non-diffusive ($\\Lambda$-effect) parts of the stress is difficult from convection simulations. Here, we present preliminary results from anisotropic homogeneous, isothermal, non-stratified turbulence simulations in which diffusive and non-diffusive effects can be studied separately. Imposing a linear shear flow on top of isotropically driven turbulence allows the study of turbulent viscosity without $\\Lambda$-effect. On the other hand, using a special form of forcing, anisotropic homogeneous turbulence can be generated. Rotation is added to study the $\\Lambda$-effect. A simple analytical closure model, based on the minimal tau-approximation (hereafter MTA, see e.g. Blackman \\& Field \\cite{BlackField2002}; Brandenburg et al.\\ \\cite{Brandea2004}), is used to compare with simulations in the cases with rotation. ", "conclusions": "\\label{sec:conclusions} Shear flow turbulence simulations show that the ratio of turbulent to molecular viscosity increases linearly up to ${\\rm Re} \\approx 30$ with $\\nu_{\\rm t}/\\nu\\approx1.5\\,\\mbox{Re}$. For the largest Reynolds number the scaling seems somewhat shallower but the present data is not yet sufficient to substantiate this. The $\\Lambda$-effect from homogeneous, anisotropic turbulence does not exhibit the puzzling features found in convection simulations. Further study is required in order to understand which of the neglected physics is responsible for the lack of these features. The MTA-closure is able to reproduce many of the qualitative aspects of the simulation results including a maximum of the horizontal stress at about $30^\\circ$ latitude, with its largest value for $\\mbox{Co}\\approx0.5$. In the model, the vertical stress can have a maximum away from the equator for $\\mbox{Co}\\ga0.2$, which is not seen in the simulations. Nevertheless, both simulations and model have the largest vertical stress for $\\mbox{Co}\\approx0.3$. However, the model generally overestimates the magnitudes of the stresses. More detailed analysis of the simulation and closure results will be presented in a future publication." }, "0710/0710.2538_arXiv.txt": { "abstract": "We analyze the possibility of delensing Cosmic Microwave Background (CMB) polarization maps using foreground weak lensing (WL) information. We build an estimator of the CMB lensing potential out of optimally combined projected potential estimators to different source redshift bins. Our estimator is most sensitive to the redshift depth of the WL survey, less so to the shape noise level. Estimators built using galaxy surveys like LSST and SNAP recover up to 80-90\\% of the potential fluctuations power at $l\\leq 100$ but only \\ax 10-20\\% of the small-angular-scale power ($l\\leq 1000$). This translates into a 30-50\\% reduction in the lensing $B$-mode power. \\par We illustrate the potential advantages of a 21-cm survey by considering a fiducial WL survey for which we take the redshift depth \\zmax and the effective angular concentration of sources \\nbar as free parameters. For a noise level of 1 $\\mu$K arcmin in the polarization map itself, as projected for a CMBPol experiment, and a beam with $\\theta_{\\rm\\scriptscriptstyle{FWHM}}$=10 arcmin, we find that going to \\zmax=20 at \\nbar=100 \\gal yields a delensing performance similar to that of a quadratic lensing potential estimator applied to small-scale CMB maps: the lensing $B$-mode contamination is reduced by almost an order of magnitude. In this case, there is also a reduction by a factor of \\ax4 in the detectability threshold of the tensor $B$-mode power. At this CMB noise level, the $B$-mode detection threshold is only $3\\times$ lower even for perfect delensing, so there is little gain from sources with $z_{max}>20$. The delensing gains are lost if the CMB beam exceeds $\\sim 20$ arcmin. The delensing gains and useful $z_{max}$ depend acutely on the CMB map noise level, but beam sizes below 10 arcmin do not help. Delensing via foreground sources does not require arcminute-resolution CMB observations, a substantial practical advantage over the use of CMB observables for delensing. ", "introduction": "\\label{I} The anisotropies in the CMB have been long recognized as a major probe for cosmology. The WMAP satellite has measured the temperature fluctuations up to $l_{max} <=$1000 and has confirmed the cosmological standard model of a power-law, flat, $\\Lambda$CDM universe. Much hope lies with CMB polarization measurements because they have the potential to unveil some of the unknowns of inflation. While some predictions of inflation (such as nearly flat space curvature, nearly scale-invariant power spectrum, and Gaussianity of the primordial fluctuations) have been confirmed by CMB and large scale structure experiments, there is another prediction which is yet to be verified. This is the existence of an almost scale-invariant spectrum of gravitational waves, whose amplitude is directly related to the energy scale of inflation. The inflationary gravitational waves are tensor perturbations to the metric and we expect them to leave a curl-like signature in the CMB polarization field, i.e. a $B$-mode pattern. Density fluctuations, arising from scalar perturbations to the metric, create a gradient-like component in the polarization field, the $E$ mode. In the linear regime, density fluctuations do not create a $B$ mode, so a detection of the latter would confirm the existence of primordial gravitational waves; it would also provide the energy scale of inflation, which we could use to distinguish between different inflationary scenarios.\\par CMB polarization measurements are difficult to carry out, primarily for two reasons. Firstly, the amplitude of the signal is very small: for example, the scalar $E$-mode power is 2-3 orders of magnitude smaller than the scalar temperature power, tensor $B$-mode much lower. Secondly, the polarization foregrounds are very poorly understood and they dominate the CMB signal at almost all relevant frequencies. Therefore, foreground removal and detector sensitivity are two of the most stifling limitations of a polarization experiment.\\par $B$-mode measurements are additionally obstructed by WL contamination, especially of the recombination signal (l$\\leq$100). Thus CMB polarization measurements are able to provide inflationary insights to the extent to which we can: 1. remove the foreground contribution to the overall signal and 2. delens the $B$ mode power and extract the tensor contribution to it. The delensing process consists of the reconstruction of the lensing potential from chosen observables. The estimated lensing potential is then used to evaluate the WL-created $B$-mode signal and to subtract it from the measured $B$-mode map. \\par In this paper we probe the ability of weak lensing (WL) galaxy surveys to delens the CMB in the absence of foregrounds. To be specific, we try to answer two questions: what attributes should a galaxy or 21~cm WL survey and also a CMB polarization mission have in order to detect the tensor $B$ mode? What is the minimum amplitude of the $B$ mode, expressed in terms of the tensor-to-scalar ratio, $r$, that can be detected using galaxy or recombination observations for delensing? We take three examples of surveys to illustrate how our estimator works: the ground-based Large Synoptic Survey Telescope (LSST \\footnote{\\tt {http://www.lsst.org/}}), the space-based Supernova Acceleration Probe (SNAP \\footnote{\\tt {http://snap.lbl.gov/}}) and a toy model mimicking recombination-era 21-cm observations that we mention in more detail in section \\S\\ref{IV}. The outline of the paper is as follows. In \\S\\ref{II} we describe the WL contamination of the tensor $B$ mode. In \\S\\ref{III} we present our minimum-variance lensing estimator. In \\S\\ref{IV} we determine the minimum detectable $r$ when we use this estimator to delens. In \\S\\ref{V} we discuss our results and draw conclusions. Let us now briefly mention similar work existing in the literature. There has been a vast and impressive amount of work on the topic of CMB delensing and $r$-detection. The great majority of this work uses the CMB observables $\\Theta, E, B$ to reconstruct the projected potential. Averaging over various quadratic combinations of the temperature field (e.g. see the work of \\citet{1999PhRvD..59l3507Z}, \\citet{1998A&A...338..767B}, \\citet{2001PhRvD..64h3005H}, \\citet{2001ApJ...557L..79H}) and the polarization field (e.g. \\citet{2000PhRvD..62d3517G}, \\citet{2002ApJ...574..566H}, \\citet{2003PhRvD..67l3507K}) has been thoroughly considered for the projected potential reconstruction. To give a quick summary: one can build minimum-variance, unbiased estimators, using certain field statistics, as shown by \\citet{2001ApJ...557L..79H}. For a post-Planck experiment (sensitivity of 0.3 $\\mu$K arcmin and beam size of 3 arcmin), \\citet{2002ApJ...574..566H} found that the most efficient of these estimators can map the potential up to l$\\leq$1000. \\citet{2002PhRvL..89a1304K} and \\citet{2002PhRvL..89a1303K} used this last estimator to predict the minimum detectable $r$ as a function of CMB experimental characteristics. There is another, more promising method for lensing potential reconstruction, based on likelihood techniques. \\citet{2003PhRvD..67d3001H} have built a maximum-likelihood estimator for the convergence field using temperature maps and have found its performance similar to that of the quadratic estimator introduced by \\citet{2001ApJ...557L..79H}. \\citet{2003PhRvD..68h3002H} found that the same maximum likelihood estimator built from polarization maps is even more effective: there is an order of magnitude reduction in the mean squared error in the lensing reconstruction compared to the quadratic estimator method, if the survey characteristics are adequate: sensitivity of 0.25 $\\mu$K arcmin and a beam size of 2 arcmin. \\citet{2005PhRvD..72l3006A} analyze the detectability of tensor $B$ modes in the presence of polarized dust emission, as a function of sky coverage; \\citet{2006JCAP...01..019V} do a study of optimal surveys for $B$-mode detection, considering both dust and synchrotron emissions. One disadvantage of the reconstruction methods presented so far is that they require high-resolution CMB maps: they use arc-minute structures of the CMB fields, to reconstruct degree-scale maps of the deflection field, as explained by \\citet{2001ApJ...557L..79H}. \\citet{2005PhRvL..95u1303S} point out that one can use non-CMB observables to delens the CMB; in this case the requirement for high angular resolution of the CMB mission can be relaxed significantly. These authors determine lower limits for the detectable $r$ using the 21 cm radiation emitted by neutral hydrogen atoms to delens the CMB. We follow a similar approach here, but employ foreground galaxies instead of 21-cm emission as the source plane for delensing. ", "conclusions": "\\label{V} In this paper we have examined the possibility of delensing $B$-mode polarization maps using galaxy WL surveys. We have proposed a weighted combination of projected potential estimators to different source redshift bins which optimally reconstructs the projected potential seen by the CMB. We have used three fiducial surveys to exemplify our estimator: LSST, SNAP and a generic survey relevant mostly to future 21-cm studies. These examples have different source redshift distribution, source density \\nbar, and sky coverage $f_{\\rm sky}$, and have enabled us to test the effect of each of these factors on the lensing potential estimator and also on its ability to reduce the WL contamination of $B$ mode maps. Using a $\\Delta \\chi^{2}$ test for the delensed $B$ mode field, we have determined the minimum value of the tensor-to-scalar ratio statistically distinguishable from 0 in optimally delensed data. Throughout this paper we have ignored the polarized foreground contamination and we have also made the assumption of Gaussianity for the delensed $B$ mode map. While this assumption may be inaccurate, especially for the cases when the efficiency of delensing is low, we expect the qualitative results of our work to hold. \\par The lensing potential estimator is sensitive mostly to the redshift depth of the WL survey and reconstructs best the large angular scale multipoles. The performance of the estimator improves continually as higher redshift source galaxies are used. In the limit \\zmax$\\rightarrow z_{\\scriptscriptstyle\\rm CMB}$ and \\nbar$\\rightarrow\\infty$, the reconstruction is perfect. An experiment like SNAP recovers \\ax 90\\% of the CMB projected potential power for $l\\leq 100$ and \\ax 20\\% for $l\\leq 1000$. The performance of LSST is a little worse, because it detects sources at an average redshift of $\\langle z \\rangle$=1, compared to $\\langle z \\rangle$=1.5 of SNAP. A box distribution (constant redshift distribution of source galaxies), provides better CMB potential estimator: if we go as far as \\zmax=20, there is an order of magnitude improvement over SNAP, for the same angular concentration of galaxies. However, in this last case, the reduction in $r_{min}$, compared to the case where no delensing is done, is only of a factor of \\ax 7 for $w^{-1/2}=0.1 \\mu$K arcmin and $\\theta_{\\rm\\scriptscriptstyle{FWHM}}$=10 arcmin. This happens because low-$l$ lensing $B$-modes are generated by beating of gravitational potential modes and $E$-modes on scales coresponding to $l$\\ax1000, where the potential estimator is less faithful. In the case of LSST and SNAP, the reduction in $r_{min}$ relative to the no delensing case is only by \\ax15\\% and \\ax50\\% for the same CMB noise level and beam. We stress that delensing with foreground weak lensing offers significant gains in $r_{min}$ (factors of 5 or more) only when three conditions are met: \\begin{enumerate} \\item the noise in the CMB map is sufficiently low, $\\sim 1\\,\\mu$K arcmin; \\item the beam size of the CMB map is $<20$ arcmin; \\item the lensing source distribution extends to $z\\sim 20$ or greater. \\end{enumerate} Delensing is not relevant to tensor mode detection if $w^{-1/2}>10 \\mu$K arcmin or if $\\theta_{\\rm\\scriptscriptstyle{FWHM}}$ \\gt$2^{\\,\\circ}$. Also, there is no advantage in lowering the beam size beyond $\\theta_{\\rm\\scriptscriptstyle{FWHM}}$=10 arcmin: for our delensing method, a 1 arcmin beam will do just as well as a 10 arcmin beam. This is perhaps the most important feature of our delensing technique, as both the quadratic estimator of \\citet{2001ApJ...557L..79H} and the maximum likelihood estimator proposed by \\citet{2003PhRvD..68h3002H} require beam sizes of 2-3 arcmin to yield their best performance. For the CMBPol mission, (which might reach a noise of 1 $\\mu$K arcmin) delensing with a box of \\zmax=20 results in $r_{min}\\approx 2\\times 10^{-5}$. If no delensing is applied, then $r_{min}\\approx 8\\times10^{-5}$, a factor of 4 worse. CMBPol detector noise keeps $r_{min}>7\\times 10^{-6}$ even with perfect delensing, so there is less incentive to acquire delensing source planes above \\zmax=20. Also for the CMBPol noise level, we have investigated the impact of $\\bar n$ and have concluded that as long as we go to high redshift ($z_{max}>10$), even with a low density of source galaxies, we can still improve $r_{min}$ by a factor of a few." }, "0710/0710.1011_arXiv.txt": { "abstract": "We present a 40 minute time series of filtergrams from the red and the blue wing of the \\halpha\\ line in an active region near the solar disk center. From these filtergrams we construct both Dopplergrams and summed ``line center'' images. Several dynamic fibrils (DFs) are identified in the summed images. The data is used to simultaneously measure the proper motion and the Doppler signals in DFs. For calibration of the Doppler signals we use spatially resolved spectrograms of a similar active region. Significant variations in the calibration constant for different solar features are observed, and only regions containing DFs have been used in order to reduce calibration errors. We find a coherent behavior of the Doppler velocity and the proper motion which clearly demonstrates that the evolution of DFs involve plasma motion. The Doppler velocities are found to be a factor 2--3 smaller than velocities derived form proper motions in the image plane. The difference can be explained by the radiative processes involved, the Doppler velocity is a result of the local atmospheric velocity weighted with the response function. As a result the Doppler velocity originates from a wide range in heights in the atmosphere. This is contrasted by the proper motion velocity which is measured from the sharply defined bright tops of the DFs and is therefore a very local velocity measure. The Doppler signal originates from well below the top of the DF. Finally we discuss how this difference together with the lacking spatial resolution of older observations have contributed to some of the confusion about the identity of DFs, spicules and mottles. ", "introduction": "\\label{sec:intro} The solar chromosphere owes its name to the reddish rim that appears above the lunar limb during solar eclipses. This reddish color mostly stems from the Balmer {\\halpha} spectral line which makes this line one of the most important chromospheric diagnostics. Due to the highly dynamic state of the chromosphere and strong NLTE effects, the line formation processes are still not yet fully understood \\citep[e.g.][]{radyn,2006Leenaarts}. This is an important shortcoming in our interpretation tools which makes \\halpha\\ observations traditionally difficult to interpret. Due to the highly fibrilar structure of the chromosphere \\citep{1908Hale}, a strong influence from magnetic fields on the chromosphere has been suspected for about a century. The most common of these fibrilar magnetic fine structures are the jet-like structures known as spicules, mottles, and dynamic fibrils (DFs). In short, spicules are traditionally observed at the limb, mottles on disk in the quiet Sun, and DFs in active regions. Whether or not these structures are manifestations of the same phenomenon viewed at different angles have been the subject of a long standing discussion \\citep[e.g.][]{1968Beckers,1992Grossman,1994Tsiropoula,1995Suematsu, 2001Christo,2007Rouppe}. One important argument against these structures being caused by the same mechanism has been the difference in the measured absolute velocities \\citep{1992Grossman}. Other authors have done direct measurements of mottles crossing the limb \\citep{2001Christo}. They also state that since both proper motions and Doppler motions are used in the comparisons, systematic errors are probably introduced. Such errors might also be amplified by the rather limited spatial resolution of some of the data sets used. The detailed analysis of DFs has accelerated in recent years \\citep[e.g.][]{2004dePontieu,2006deWijn,2006Hansteen,2007dePontieu, 2007Lars} due to major advances in both observational techniques and simulation efforts. One of the main conclusions from these studies is that DFs are driven by magneto-acoustic shocks caused by p-mode oscillations and convective flows leaking into the chromosphere. In a recent study, \\citet[][from now on paper 1]{2007bLangangen} presented spectroscopic analysis of DFs as seen in one of the Ca~{\\small{II}}~IR lines. Numerical analysis of the line formation process showed a much lower DF velocity derived from Doppler measurements as compared to the proper motion velocity. This was found to be due to both the low formation height and the extensive width of the contribution function of the Ca~{\\small{II}}~IR line. Furthermore, the DFs analyzed in paper 1 showed mass motion, thus ruling out any ionization/temperature wave as explanation model for DFs \\citep[e.g.][and references therein]{2000Sterling}. With the advantage of well sampled spectral line profiles, the number of analysed DFs in paper~1 was rather modest due to the limited spatial coverage of the spectrograph slit. With the current data we exploit the wide spatial coverage of a tunable filter instrument, at the expense of limited spectral resolution. \\begin{figure*}[!ht] \\includegraphics[width=\\textwidth]{f1.eps} \\caption{ Field of view (FOV) for one wideband image (left) and the corresponding sum of the two narrow band filtergrams (right). In the narrow band image several fibrilar structures are seen, and some DF axes are marked (solid white lines) for illustrative purposes. } \\label{plotone} \\end{figure*} In this paper we add to the understanding of these jet like structures by analysis of Dopplergrams obtained in an active region close to the disk center, hence the observed jet structures are commonly known as DFs. In \\S\\ref{sec:obs} we describe the observational program and the instrumentation. The data reduction and the calibration method is explained in \\S\\ref{sec:datareduction}. In \\S\\ref{sec:Obsres} we present the results of our measurements. We discuss our results in \\S\\ref{Disc} and finally we summarize the results in \\S\\ref{Sum}. ", "conclusions": "\\label{Disc} The correlation between the maximum velocities and decelerations found from the proper motion measurements is similar to the correlation found by \\citet{2006Hansteen} and \\citet{2007dePontieu}. This is illustrated in Fig.~\\ref{plotseven} by the grey-scaled cloud shown in the background. This correlation between the deceleration and the maximum velocity is known to be the signature of shock waves being the driving mechanism of DFs \\citep{2006Hansteen,2007dePontieu,2007Lars}. Further support for this driving mechanism comes from the fact that we find a coherent behavior between the evolution of the Doppler signal and the proper motion for a large fraction of the DFs. This is a strong indication that there is actual plasma motion occurring during the life time of DFs. This supports the findings of paper~1, but based on a much larger sample. The Doppler measurements show a similar correlation as for the proper motion, but with much lower absolute values for both the decelerations and maximum velocities. One possible explanation for these lower values could be high inclination angles of the DF trajectories with the LOS. One could expect to be able to derive the full trajectory vector by combining the two measured deceleration components. This naive method would give very high inclination angles, typically $75^\\circ$. We know, however, that this can not be the true inclination angle, since the Doppler velocity is a result of the local atmospheric velocity weighted with the response function to velocity over an extended height. In contrast, the measured proper motion is very local due to the high contrast boundary between the top of the fibril and the surroundings. Combining these two measurements leads to highly overestimated inclination angles. The difference in absolute values must be considered in the context of the results from Paper~1. The lower Doppler velocities found from the Ca~{\\small{II}}~IR line was explained by a combination of lower formation height and extended formation range. This is probably also the case for the \\halpha\\ line, but the formation height usually extends over a larger height range as compared to the Ca~{\\small{II}}~IR line. The lacking Doppler shifts in ${\\sim}15$\\% of the DFs can either be caused by very high inclination angles, or their driving mechanism is fundamentally different and the evolution of these DFs is not a result of mass motion. We believe that high inclination angles combined with the uncertainties in the measurements is a more plausible explanation for the lacking Doppler signals. The identification method of DFs introduces a bias toward the more inclined DFs. There are a number of suggestive cases where DFs are visible in the Dopplergrams but no clear signature can be seen in the corresponding intensity images. We refrain from measuring these DFs since this will complicate a comparison with other data sets. Furthermore, the identification of these DFs is not objective and we expect the measurement errors to be unacceptably high since low inclination angles would lead to potentially high Doppler velocities. Due to the lacking spectral sampling this could lead to strong saturation effects in the measured Doppler velocities. \\subsection{Spicules, mottles and fibrils} The identification of the disk counterpart of spicules was already an important question forty years ago \\citep[e.g][]{1968Beckers}. One of the main problems was to reconcile the velocities measured in spicules with those measured in mottles. This problem was also the main concern of \\citet{1992Grossman}. They conclude that since the velocities are much larger in spicules than in mottles, the two could not be the same structure seen at different viewing angles. They, however, admit that the seeing might impair their results if the structures observed were smaller than $1\\arcsec{}$. Later studies of mottles and spicule properties lead to the conclusion that spicules and mottles are in fact the same feature seen at different angles \\citep{1993Tsiropoula,1994Tsiropoula}. \\citet{1994Tsiropoula} showed, using cloud modelling, that the proper motion and the cloud velocities were consistent. In a more recent work, \\citet{2001Christo} use a limb darkening correction method to directly observe mottles crossing the limb. They argue that the main reason for earlier confusion is caused by the lacking spatial resolution of the observations. In our study it is clear that the Doppler signal originates from spatially resolved structures. The excellent quality of the observations largely removes the errors due to lacking spatial resolution. Assuming reasonable inclination angles, i.e. the mean angles being not very large nor very small, we can conclude that the Doppler velocities are typically a factor of $\\sim 2$--$3$ smaller than the corresponding proper motion. A similar, but larger difference was reported by \\citet{1994Tsiropoula}, we believe that this difference can be attributed to worse spatial resolution. As discussed above, radiative transfer processes are the fundamental reason for the Doppler velocities being lower than the proper motion. We argue that the fundamental differences between Doppler and proper motion velocities that we find for DFs, also are valid for similar measurements on spicules and mottles. Besides the arguments of lacking spatial resolution, we believe that this difference was an important contributor to the earlier confusion about the unification of mottles and spicules. Similar work on spatially resolved limb spicules is needed to finally settle this discussion." }, "0710/0710.4919_arXiv.txt": { "abstract": "Quantum gravity may remove classical space-time singularities and thus reveal what a universe at and before the big bang could be like. In loop quantum cosmology, an exactly solvable model is available which allows one to address precise dynamical coherent states and their evolution in such a setting. It is shown here that quantum fluctuations before the big bang are generically unrelated to those after the big bang. A reliable determination of pre-big bang quantum fluctuations would require exceedingly precise observations. ", "introduction": "Cosmology, as the physics of the universe as a whole, places special limitations on scientific statements to be reasonably inferred from observations. On a general basis, this was discussed in \\cite{Uncertain} who noted that some general properties seem to arise automatically during cosmological evolution. Their verification for our universe then does not reveal anything about its possibly more complicated past. For instance, isotropization \\cite{Isotropize} demonstrates that the observation of a nearly isotropic present universe does not present much useful information on an initial state at much earlier times. Similarly, one may add decoherence to the list which can be seen as doing the same for the observation of a nearly classical present universe: Most initial states would decohere and arrive at a semiclassical state; thus its observation does not rule out stronger quantum behavior at earlier stages. Given this, it is surprising to see recent discussions about a universe before the big bang in several approaches to quantum gravity. Not just statements about the possibility of the universe having existed before the big bang are being made, which could in principle be inferred from a general analysis of equations of motion, but even assumptions on its classicality or claims about the form of its state such as its fluctuations at those times are put forward. Even if this may be possible theoretically, it raises the question how much one can really learn about a pre-big bang universe from observations. A solvable model of quantum cosmology is used here to draw cosmological conclusions within this model about the possible nature of the big bang and a universe preceding it. The viewpoint followed has been spelled out in \\cite{BeforeBB} where also some of the results have already been reported. This model does show a bounce at small volume instead of the classical singularity present in solutions of general relativity. We therefore use it to shed light on the general questions posed above. The promotion of this specific model is not intended as a statement that its bounce would be generic for quantum gravity even within the same framework. In fact, the conclusion of the bounce in this and related models available so far is based on several specific properties which prevent a generic statement about this form of singularity removal. We rather take the following viewpoint: Assume there is a theoretical description of a bouncing universe; what implications can be derived for its pre-bounce state? The specific model used is distinguished by a key fact which makes it suitable for addressing such a general question: As a solvable model, the system displays properties of its dynamical coherent states which, to some degree, are comparable to those of the harmonic oscillator. There is no back-reaction of quantum fluctuations or higher moments of a state on the time evolution of its expectation values. Their trajectory is thus independent of the spreading of a state if it occurs, a property which is well known for the harmonic oscillator (for instance, as a simple consequence of the Ehrenfest theorem). This behavior is certainly special compared to other quantum systems, but the precise derivation of properties of dynamical coherent states it allows are nevertheless important. For instance, a large part of theoretical quantum optics is devoted to a computation of fluctuations in squeezed coherent states of the harmonic oscillator. Similarly, the solvable model studied here allows detailed calculations of its dynamical coherent states which are of interest for quantum cosmology. In particular, the solvable model we will be using eliminates the classical big bang singularity by quantum effects. Dynamical coherent states highlight the behavior of fluctuations of the state of the universe before and after the big bang. Care is, however, needed for the physical interpretation of the results. While quantum optics allows one to prepare a desired state and perform measurements on it, quantum cosmology has to make use of the one universe state that is given to us. Unlike quantum optics, where states can be prepared to be close to harmonic oscillator states, we cannot realize states of the solvable quantum cosmological model. The real universe is certainly very different from anything solvable, and so the availability of a certain feature or numerical result in a solvable model is unlikely to be related to a realistic property of the universe. Thus, as spelled out in \\cite{BeforeBB}, we focus on pessimism in our analysis of the solvable model: the {\\em in}ability of making certain predictions even in a fully controlled model is likely to be a reliable statement much more generally; adding complications of a real universe would only make those predictions even more difficult. For solvable systems of this kind it is easier by far to solve equations of motion for expectation values and fluctuations directly, rather than taking the detour of a specifically represented wave function. Properties of coherent states are then determined by selecting solutions of fluctuations which saturate uncertainty relations. We will first illustrate this procedure for the harmonic oscillator with an emphasis on squeezed states. We present this brief review in section \\ref{Squeezed} intended as an introduction of the methods then used in quantum cosmology. The main part of this paper is an application of those methods to the solvable system of quantum cosmology, as well as a physical discussion. Instead of different cycles of a harmonic oscillator, in quantum cosmology we will be dealing with the pre- and post-big bang phase of a universe. Just as fluctuations in a squeezed harmonic oscillator state can oscillate during the cycles, fluctuations in quantum cosmology may change from one phase to the next. Our main concern will be the reliability of predictions about the precise state of the universe before the big bang, based on knowledge we can achieve after the big bang. ", "conclusions": "" }, "0710/0710.3364_arXiv.txt": { "abstract": "The focusing of the radiation generated by a polarization current with a superluminally rotating distribution pattern is of a higher order in the plane of rotation than in other directions. Consequently, our previously published asymptotic approximation to the value of this field outside the equatorial plane breaks down as the line of sight approaches a direction normal to the rotation axis, \\ie is nonuniform with respect to the polar angle. Here we employ an alternative asymptotic expansion to show that, though having a rate of decay with frequency ($\\mu$) that is by a factor of order $\\mu^{2/3}$ slower, the equatorial radiation field has the same dependence on distance as the nonspherically decaying component of the generated field in other directions: it, too, diminishes as the inverse square root of the distance from its source. We also briefly discuss the relevance of these results to the giant pulses received from pulsars: the focused, nonspherically decaying pulses that arise from a superluminal polarization current in a highly magnetized plasma have a power-law spectrum (\\ie a flux density $S\\propto\\mu^\\alpha$) whose index ($\\alpha$) is given by one of the values $-2/3$, $-2$, $-8/3$, or $-4$. ", "introduction": "Radiation by polarization currents whose distribution patterns move faster than light {\\it in vacuo} has been the subject of several theoretical and experimental studies in recent years \\cite{BessarabAV:FasEsi,ArdavanA:Exponr,BessarabAV:Expser,BolotovskiiBM:Radssv,% BolotovskiiBM:Radbcm,ArdavanH:Genfnd,ArdavanH:Speapc,ArdavanH:Morph}. When the motion of its source is accelerated, this radiation exhibits features that are not shared by any other known emission. In particular, the radiation from a rotating superluminal source consists, in certain directions, of a collection of subbeams whose azimuthal and polar widths narrow (as $R\\subP^{-3}$ and $R\\subP^{-1}$, respectively) with distance $R\\subP$ from the source \\cite{ArdavanH:Morph}. Being composed of tightly focused wave packets that are constantly dispersed and reconstructed out of other waves, these subbeams neither diffract nor decay in the same way as conventional radiation beams. The field strength within each subbeam diminishes as $R\\subP^{-1/2}$, instead of $R\\subP^{-1}$, with increasing $R\\subP$ \\cite{ArdavanH:Genfnd,ArdavanH:Speapc,ArdavanH:Morph}. In earlier treatments \\cite{ArdavanH:Speapc,ArdavanH:Morph}, we evaluated the field of a superluminally rotating extended source by superposing the fields of its constituent volume elements, \\ie by convolving its density with the familiar Li\\'enard-Wiechert field of a rotating point source. This Li\\'enard-Wiechert field is described by an expression essentially identical to that which is encountered in the analysis of synchrotron radiation, except that its value at any given observation time receives contributions from more than one retarded time. The multivalued nature of the retarded time gives rise to the formation of caustics. The wave fronts emitted by each constituent volume element of a superluminally moving accelerated source possess a cusped envelope on which the field is infinitely strong (see Figs.~1 and 4 of Ref.~\\cite{ArdavanH:Morph}). Correspondingly, the Green's function for the problem is nonintegrably singular for those source elements that approach the observer along the radiation direction with the speed of light and zero acceleration at the retarded time (see Fig.~3 of Ref.~\\cite{ArdavanH:Morph}): the cusp of the envelope of wave fronts emanating from each such element is a spiraling curve extending into the far zone that passes through the position of the observer. When the source oscillates at the same time as rotating, the Hadamard finite part of the divergent integral that results from convolving the Green's function with the source density has a rapidly oscillating kernel for a far-field observation point. The stationary points of the phase of this kernel turn out to have different orders depending on whether the observer is located in or out of the equatorial plane. To reduce the complications posed by the higher-order stationary points of this phase, we restricted the asymptotic evaluation of the radiation integral thus obtained in Refs.~\\cite{ArdavanH:Speapc,ArdavanH:Morph} to observation points outside the plane of rotation, \\ie to spherical polar angles $\\theta\\subP$ that do not equal $\\pi/2$. The purpose of this paper is to evaluate the field of a superluminally rotating extended source also for the smaller class of observers at polar coordinate $\\theta\\subP=\\pi/2$ and to obtain, thereby, a more global description of the nonspherically decaying radiation that is generated by such a source. The asymptotic expansion presented in Refs.~\\cite{ArdavanH:Speapc,ArdavanH:Morph} breaks down in the case of a subbeam that is perpendicular to the rotation axis because there is a higher-order focusing associated with the waves emitted by those source elements whose actual speeds (rather than the line-of-sight components of their speeds) equal the speed of light as they approach the observer with zero acceleration. Here, we present a brief account of the background material on the radiation field of a rotating superluminal source in Section 2, and the asymptotic evaluation of this field for an equatorial observer in Section 3. In Section 4, we give a description of the spectral properties of the nonspherically decaying component of this radiation in the light of the present results and those obtained in Refs.~\\cite{ArdavanH:Speapc,ArdavanH:Morph}, and discuss the relevance of these properties to pulsar observations. ", "conclusions": "" }, "0710/0710.2542_arXiv.txt": { "abstract": "{INTEGRAL, the European Space Agency's $\\gamma$-ray observatory, tripled the number of super-giant high-mass X-ray binaries (sgHMXB) known in the Galaxy by revealing absorbed and fast transient (SFXT) systems.} {In these sources, quantitative constraints on the wind clumping of the massive stars could be obtained from the study of the hard X-ray variability of the compact accreting object.} {Hard X-ray flares and quiescent emission of SFXT systems have been characterized and used to derive wind clump parameters.} {A large fraction of the hard X-ray emission is emitted in the form of flares with a typical duration of 3 ks, frequency of 7 days and luminosity of $10^{36}$ erg/s. Such flares are most probably emitted by the interaction of a compact object orbiting at $\\sim10~R_*$ with wind clumps ($10^{22-23}$ g) representing a large fraction of the stellar mass-loss rate. The density ratio between the clumps and the inter-clump medium is $10^{2-4}$ in SFXT systems. } {The parameters of the clumps and of the inter-clump medium, derived from the SFXT flaring behavior, are in good agreement with macro-clumping scenario and line driven instability simulations. SFXT have probably a larger orbital radius than classical sgHMXB.} ", "introduction": "\\label{sec:intro} Stellar winds have profound implications for the evolution of massive stars, on the chemical evolution of the Universe and as source of energy and momentum in the interstellar medium. Photons emitted by massive stars directly transfer momentum to the stellar wind through absorption in numerous Doppler shifted optically thick spectral lines \\citep{cak1975} and drive the wind to highly supersonic speeds. Such line driven stellar winds are very unstable \\citep{ocr1988,feldmeier1995}. The collisions between high speed and low density material with slower gas trigger strong shocks, form high density contrasts and wind clumps and lead to thermal X-ray emission. There are multiple observational lines of evidence for wind clumping in massive stars from wind line profiles \\citep{Bouret2005,Fullerton2006}, discrete absorption components \\citep{PrinjaHowarth1986}, variable line profiles \\citep{Lepine1999,Markova2005}, polarimetry \\citep{Lupie1987, Davies2007}, X-rays continuum \\citep{CassinelliOlson1979} and line emission \\citep{OskinovaFeldmeierHamann2006}. Wind clumping has an important effect on mass-loss rate diagnostics that depends on the square of the wind density. It leads to a reduction of the stellar mass-loss rate estimates by a factor $f_V^{-0.5}$ where $f_V$ is the clump volume filling factor \\citep{Abbott1981,Fullerton2006}. A reduction by a factor $>3$ becomes problematic for massive star evolution \\citep{Hirschi2007}. Optically thick clumps may however help to reconcile wind clumping and usual mass-loss rates by reducing spectral features and free-free emission \\citep{OskinovaHamannFeldmeier2007}. Indirect measures of the structure of massive star winds are possible in binary systems through the analysis of the interaction between the stellar wind and the companion or its stellar wind. In colliding wind binaries, \\cite{Schild2004} and \\cite{Pollock2005} have pointed out that the observed X-ray column densities are much lower than expected from smooth winds and that clumped winds are the probable solution \\citep[see also][]{pittard2007}. In wind-fed accreting High-Mass X-Ray Binaries (HMXB), clumping has been invoked to explain the orbital variability of X-ray line profiles emitted by the photo-ionized wind \\citep{Sako2003,Vandermeer2005}. In this paper we aim at studying the clumping of stellar wind in HMXB using the hard X-ray variability observed by the IBIS/ISGRI instrument \\citep{ubertini03AA,lebrun03AA} on board the INTErnational Gamma-Ray Astrophysics Laboratory \\citep[INTEGRAL,][]{winkler03AA} which could be used to probe clump parameters and density contrasts in the stellar wind. This study follows a first paper \\citep{Leyder2007} interpreting the flaring behavior of \\object{IGR J08408$-$4503} in term of wind clumping. Classical wind-fed super-giant HMXB (sgHMXB) are made of a compact object orbiting within few stellar radii from a super-giant companion. Recently INTEGRAL almost tripled the number of sgHMXB systems known in the Galaxy and revealed a much more complex picture with two additional families of sources: the highly absorbed persistent systems \\citep{walter04esasp,walter2006} and the super-giant fast X-ray transient (SFXT) systems \\citep{Negueruela2006}. The highly absorbed systems have orbital and spin periods similar to those of classical Roche-lobe underflow sgHMXB, however the absorbing column densities are much higher than observed on average in classical systems \\citep{walter2006}. The fast transient systems are characterized by fast outbursts, by a low quiescent luminosity and by super-giant OB companions \\citep{Sguera2006,Negueruela2007}. The sample of sources and the INTEGRAL data analysis are described in Sect. 2. Their hard X-ray variability is discussed in the context of clumpy stellar winds in Sect. 3. Finally Sect. 4 summarizes our principal conclusions. ", "conclusions": "INTEGRAL tripled the number of super-giant HMXB systems known in the Galaxy and revealed two new populations: the absorbed and the fast transient (SFXT) systems. The typical hard X-ray variability factor is $\\lesssim 20$ in classical and absorbed systems and $\\gtrsim 100$ in SFXT. We have also identified some ``intermediate'' systems with smaller variability factors that could be either SFXT or classical systems. The SFXT behavior is best explained by the interaction between the accreting compact object and a clumpy stellar wind \\citep{IntZand2005,Leyder2007}. Using the hard X-ray variability observed by INTEGRAL in a sample of SFXT we have derived typical wind clump parameters. The compact object orbital radius are probably relatively large ($10~R_*$) and the clumps which generate most of the hard X-ray emission have a size of a few tenth of $R_*$. The clump mass is of the order or $10^{22-23}~\\rm{g}$ (for a column density of $10^{22-23}~\\rm{cm}^{-2}$) and the corresponding mass-loss rate is $10^{-(5-6)}~\\rm{M_{\\odot}/y}$. At the orbital radius, the clump separation is of the order of $R_*$ and their volume filling factor is $0.02$. Depending how the clump density varies with radius, the average volume filling factor could be as large as 0.1. These parameters are in good agreement with the macro clumping scenario proposed by \\cite{OskinovaHamannFeldmeier2007}. The observed ratio between the flare and quiescent count rates indicate density ratios between the clumps and the inter-clump medium which vary between 15 to 50 in ``Intermediate'' systems and $10^{2-4}$ in SFXT. Such ratios and the observed clump densities are in reasonable agreement with the predictions of line driven instabilities at large radii \\citep{Runacres2005}. The main difference between classical sgHMXB and SFXT could be the orbital radius of the compact object. At small orbital radius ($R_{orb}\\sim2~R_*$) the systems are persistent and luminous. At larger radius and if wind clumping takes place the fast transient SFXT behavior is observed." }, "0710/0710.0547_arXiv.txt": { "abstract": "We present new $H$-band echelle spectra, obtained with the NIRSPEC spectrograph at Keck II, for the massive star cluster ``B'' in the nearby dwarf irregular galaxy NGC~1569. From spectral synthesis and equivalent width measurements we obtain abundances and abundance patterns. We derive an Fe abundance of [Fe/H]=$-0.63\\pm0.08$, a super-solar [$\\alpha$/Fe] abundance ratio of $+0.31\\pm0.09$, and an O abundance of [O/H]=$-0.29\\pm0.07$. We also measure a low $\\rm ^{12}C/^{13}C\\approx 5\\pm1$ isotopic ratio. Using archival imaging from the Advanced Camera for Surveys on board HST, we construct a colour-magnitude diagram (CMD) for the cluster in which we identify about 60 red supergiant (RSG) stars, consistent with the strong RSG features seen in the $H$-band spectrum. The mean effective temperature of these RSGs, derived from their observed colours and weighted by their estimated $H$-band luminosities, is 3790 K, in excellent agreement with our spectroscopic estimate of $T_{\\rm eff} = 3800\\pm200$ K. From the CMD we derive an age of 15--25 Myr, slightly older than previous estimates based on integrated broad-band colours. We derive a radial velocity of $\\langle v_r \\rangle=-78\\pm 3$ km/s and a velocity dispersion of $9.6\\pm0.3$ km/s. In combination with an estimate of the half-light radius of $0\\farcs20\\pm0\\farcs05$ from the HST data, this leads to a dynamical mass of $(4.4\\pm1.1)\\times10^5$ M$_\\odot$. The dynamical mass agrees very well with the mass predicted by simple stellar population models for a cluster of this age and luminosity, assuming a normal stellar IMF. The cluster core radius appears smaller at longer wavelengths, as has previously been found in other extragalactic young star clusters. ", "introduction": "Massive star clusters are potentially useful test particles for constraining the evolutionary histories of their host galaxies. They can remain observable for the entire lifetime of a galaxy (as illustrated by the old \\emph{globular clusters} which surround every major galaxy), and they are bright enough to be studied in detail well beyond the Local Group. The internal velocity broadening of their spectra is typically only a few km/s, making detailed abundance analysis at high spectral resolution feasible. In a previous paper we have taken a first step towards exploiting this potential by analysing $H$ and $K$-band spectra of a young massive star cluster (YMC) in the nearby spiral galaxy NGC~6946 \\citep{lar06}. Here we apply a similar analysis to one of the young star clusters in the nearby (post-) starburst galaxy NGC~1569. NGC~1569 was one of the first galaxies in which the presence of exceptionally bright, young star clusters was suspected. A spectrum of one of the two bright ``stellar condensations'' in NGC~1569 was obtained already by \\citet{mayall35}, although the true nature of these objects was probably first discussed in detail by \\citet{as85}. They found that the spectra are of composite nature, and also noted that observations with the Hubble Space Telescope (HST) would definitively settle the issue whether or not these objects are really ``super star clusters''. Indeed, pre-refurbishment mission HST observations by \\citet{oco94} settled the issue by showing that both objects are extended, with half-light radii of about two pc. Based on high-dispersion spectroscopy from the HIRES spectrograph on the Keck~I telescope, a dynamical mass of $\\approx 10^6$ M$_\\odot$ was soon after estimated for the brighter of the two objects, NGC~1569-A \\citep{hf96,stern98}. The clusters in NGC~1569 are prime candidates for abundance analysis in the near-infrared where they are very bright. The IR holds a particular advantage over optical studies for NGC~1569 due to the large amount of foreground extinction. \\citet{gg02} found the integrated $H$-band spectra of both clusters A and B to be well approximated by red supergiant templates. Although NGC~1569-A is the brighter of the two, it is actually a binary cluster itself with some evidence for a (small) age difference between the two components \\citep{guido97,maoz01,origlia2001}. Whether or not the two components are physically connected or the result of a chance projection is unknown. In this paper we concentrate on NGC~1569-B. In addition to providing insight into the histories of their host galaxies, young star clusters with masses in excess of $10^5$ M$_\\odot$ also offer an excellent opportunity to study large samples of coeval massive stars. Such clusters are rare in the Milky Way and even in the Local Group, so it is necessary to extend the search to a larger volume. The most massive known young star clusters in the Milky Way disk are Westerlund 1 \\citep{clark05} and an object discovered in the 2MASS survey \\citep{figer06}, both of which may have masses approaching $10^5$ M$_\\odot$. Young clusters of similar masses are found in the Large Magellanic Cloud \\citep{vdb99}, but these pale in comparison with objects like NGC~1569-A and NGC~1569-B. We have obtained new NIRSPEC $H$-band spectra of NGC~1569-B, optimised for abundance analysis, and we additionally present photometry for \\emph{individual} stars in the cluster derived from archival observations with the high resolution channel (HRC) of the Advanced Camera for Surveys (ACS) on HST. The HST data provide an independent verification of the stellar parameters (notably $T_{\\rm eff}$ and $\\log g$) derived from the spectral analysis, and also allow us to compare the observed colour-magnitude diagram (CMD) with standard isochrones. Finally, we use new measurements of the structural parameters and velocity dispersion to derive the cluster mass and mass-to-light ratio and compare with predictions by simple stellar population (SSP) models. We begin by briefly describing the data in \\S\\ref{sec:data}. The analysis and our main results then follow in \\S\\ref{sec:results}, where we first discuss the near-infrared spectroscopy and the abundance analysis (\\S\\ref{sec:abundance}). We then proceed to construct a CMD from the HST imaging (\\S\\ref{sec:cmd}) from which we derive stellar parameters for the red supergiants (RSGs) that are compared against the spectroscopic results (\\S\\ref{sec:speccmp}) and theoretical isochrones (\\S\\ref{sec:interp}). In \\S\\ref{sec:virmass} we measure structural parameters for NGC~1569-B and combine these with velocity dispersion measurements to derive a dynamical mass estimate. Finally, some additional discussion and a summary are given in \\S\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} We have presented a detailed investigation of the stellar content and other properties of the massive star cluster 'B' in NGC~1569. Using new high S/N $H$-band echelle spectroscopy from the NIRSPEC spectrograph on Keck~II, we have carried out abundance analysis of red supergiant stars in the cluster. We find an iron abundance of [Fe/H] = $-0.63\\pm0.08$, close to that of SMC field stars. Our estimate of the oxygen abundance, [O/H] = $-0.29\\pm 0.07$, is about a factor of two higher than that derived for H{\\sc ii} regions \\citep{sdk94,dev97,ks97}, and the resulting super-solar [O/Fe] abundance is unlike that observed in the Magellanic Clouds and young stellar populations in the Milky Way. However, according to our measurements NGC~1569-B also differs from these galaxies by having a higher [$\\alpha$/Fe] ratio. The difference between our [O/H] measurement for NGC~1569-B and the O abundance derived for H{\\sc ii} regions is greater than the formal uncertainties on either value, and our oxygen abundance is higher than the nebular abundances in most dwarf galaxies \\citep[e.g.][]{vad07}. This raises the question whether one (or both) methods suffers from possible systematic errors, or if there might be a genuine difference between cluster and H{\\sc ii} region abundances in NGC~1569 -- perhaps due to an ab initio oxygen enhancement of the gas out of which the cluster formed. Since oxygen is the only element in common between the cluster and H{\\sc ii} region data, a detailed comparison of the abundance patterns is unfortunately not possible. However, it is of interest to note that the very strong Wolf-Rayet features in the spectrum of cluster A also suggest a high metallicity \\citep{maoz01}. NGC~1569 is a popular target for testing models of chemical evolution, and it is generally recognised that the strong galactic wind observed in the galaxy must play an important role. \\citet{martin02} observed a super-solar $[\\alpha/{\\rm Fe}]$ ratio for the wind, as we do for NGC~1569-B. However, it is unclear to what extent the two are related as material ejected in the outflow may not participate in star formation, and in any case not before it has had time to cool down and fall back into the galaxy. The chemical evolution of an NGC~1569-type galaxy has been modelled in detail by \\citet{rec06}, assuming a variety of bursty star formation histories. Their chemo-dynamical simulations show that elements produced in the last burst of star formation generally do not get mixed with the ISM in the galaxy, but instead get injected into the hot gas phase. None of their models produce an O abundance as high as the one observed by us, and the model that comes closest ($12 + \\log$ (O/H) $\\sim 8.4$) severely underpredicts the N/O abundance of the H{\\sc ii} regions observed by \\citet{ks97}. Generally, the best fits to the H{\\sc ii} region abundances are obtained for a ``gasping'' star formation history, but these models all predict a maximum $12 + \\log$ (O/H) $\\approx$ 8.1. Similarly, the O abundance predicted by the models of \\citet{romano06} for NGC~1569 never exceeds $12 + \\log$ (O/H) = 8.41. We speculate that the galactic wind which is responsible for removing metals may be less efficient in doing so near the bottom of the potential well where the massive clusters formed. Unfortunately, none of the models include predictions for the wide range of other elements we are observing here. The Small Magellanic Cloud presents an interesting comparison case. Early spectroscopic studies and Str{\\\"o}mgren photometry indicated that the young cluster NGC~330 is about 0.5 dex more metal-poor than the surrounding field \\citep[and references therein]{gr92}. High-dispersion spectroscopy by \\citet{hill99} showed a smaller and only marginally significant difference between the field stars and NGC~330, although still in the sense that the cluster is more metal-poor than the field ([Fe/H] = $-0.82\\pm0.10$ vs.\\ [Fe/H] = $-0.69\\pm0.11$). The [O/Fe] abundance ratios in these stars were all found to be sub-solar ([O/Fe] $\\approx$ $-$0.15 to $-$0.3 dex), as in the LMC and in young Galactic supergiants \\citep{hbs97}. \\citet{hill99} also found the $\\alpha$-elements abundances (Mg, Ca, Ti) relative to Fe to be around Solar for NGC~330. \\citet{gw99} found slightly lower iron abundances of [Fe/H] = $-0.94\\pm0.02$ for K supergiants in NGC~330 than \\citet{hill99} and an [O/Fe] ratio closer to Solar. They derived somewhat enhanced [Ca/Fe], [Si/Fe] and [Mg/Fe] ratios ($+0.18$, $+0.32$ and $+0.11$), in contrast to sub-solar ratios derived for B stars. While there is some evidence for a difference in metallicity between NGC~330 and the SMC field, the situation there seems to contrast with the case of NGC~1569-B where we find a \\emph{higher} oxygen abundance of the cluster compared to the H{\\sc ii} regions and an \\emph{enhanced} [O/Fe] ratio. Our mean [$\\alpha$/Fe] = $+0.31\\pm0.09$ is also higher than that derived for the stars in NGC~330. One significant difference may be the mass of NGC~1569-B, which is probably an order of magnitude greater than that of NGC~330 \\citep{fb80}. The spectral analysis returns a mean $T_{\\rm eff} = 3800\\pm200$ K and $\\log g \\approx 0.0$ for the RSGs in NGC~1569-B. We have checked these values using resolved photometry of the RSGs in the cluster, derived from archival HST/ACS images. About 60 RSGs are easily identifiable in the colour-magnitude diagram, and from their $m_{\\rm F555W} - m_{\\rm F814W}$ colours we derive a mean effective temperature of $\\langle T_{\\rm eff}\\rangle = 3850$ K and $\\log g = 0.1$. An even closer match to the spectroscopic $T_{\\rm eff}$ is obtained if we weigh the $T_{\\rm eff}$ estimates for the individual stars by their $H$-band luminosities, in which case we get $\\langle T_{\\rm eff} \\rangle = 3790$ K. We have compared the CMD with $Z=0.004$ and $Z=0.008$ isochrones from the Padua and Geneva groups. Both sets of models provide the best match to the observed colours and magnitudes of the RSGs for $Z=0.008$ and ages of 15--25 Myrs, but no isochrone can reproduce the observed CMD in detail. For these sub-solar metallicities, the models predict higher (by up to a few 100 K) mean effective temperatures $\\langle T_{\\rm eff} \\rangle$ for the RSGs than observed. Since the RSGs generally become cooler with increasing metallicity, models of higher metallicity are favoured, but models of Solar metallicity would be required to match the observed colours. For such models, however, the \\emph{blue} supergiants become much too cool to match the observations. The observed ratio of blue to red supergiants (BSG/RSG = $0.39\\pm0.10$) is also significantly lower than most of the model predictions. Although we are detecting an impressive number of RSGs, it should be noted that since we are only resolving stars located well outside the half-light radius, the actual number of RSGs in NGC~1569-B is likely to be more than a factor of two greater than the number we have detected. We derive a velocity dispersion of $9.6\\pm0.3$ km/s from the integrated spectrum of NGC~1569-B, somewhat higher than the value of 7.5 km/s of \\citet{gg02}. Combining this with an estimate of the cluster half-light radius of $0\\farcs20\\pm0\\farcs05$ measured on the ACS images, we obtain a dynamical mass of $(4.4\\pm1.1)\\times10^5 \\left(\\frac{D}{2.2 {\\rm Mpc}}\\right) M_\\odot$. This is in excellent agreement with the mass predicted by simple stellar population models for a cluster of this age and luminosity and a standard IMF. A correction for mass segregation would bring the two estimates into even closer agreement. Our analysis of structural parameters reveals a decrease in core radius with wavelength for NGC~1569-B. This trend is consistent with the colour gradients observed by \\citet{hunter00} and by us in the HRC data when variations in the PSF are taken into account. A similar colour gradient was observed in an YMC in NGC~6946 \\citep{laretal01}. The correlation between core radius and wavelength is also in the same sense as that found for M82-F by \\citet{mgv05} and may be an indication that mass segregation (primordial or dynamical) is present in NGC~1569-B. However, this needs to be confirmed by a more detailed analysis, including a modelling of how mass segregation would translate into observables such as core- and half-light radius. As a final note, we find it fascinating to contemplate that only a couple of decades ago, it was still uncertain whether NGC~1569-A and NGC~1569-B were in fact star clusters in NGC~1569 or mere foreground stars. Today, the Advanced Camera for Surveys on HST has made it possible to not only settle this question definitively, but to study the individual stars that make up these clusters." }, "0710/0710.0637_arXiv.txt": { "abstract": "We combine high-resolution images in four optical/infra-red bands, obtained with the laser guide star adaptive optics system on the Keck Telescope and with the Hubble Space Telescope, to study the gravitational lens system \\lensname (lens redshift~\\zdvalue, source redshift~\\zsvalue). We show that (under favorable observing conditions) ground-based images are comparable to those obtained with HST in terms of precision in the determination of the parameters of both the lens mass distribution and the background source. We also quantify the systematic errors associated with both the incomplete knowledge of the PSF, and the uncertain process of lens galaxy light removal, and find that similar accuracy can be achieved with Keck LGSAO as with HST. We then exploit this well-calibrated combination of optical and gravitational telescopes to perform a multi-wavelength study of the source galaxy at $0\\farcs01$ effective resolution. We find the \\sersic index to be indicative of a disk-like object, but the measured half-light radius (\\re=\\revalue) and stellar mass (\\stellarmass=\\stellarmassvalue) place it more than three sigma away from the local disk size-mass relation. The \\lensname source has the characteristics of the most compact faint blue galaxies studied, and has comparable size and mass to dwarf early-type galaxies in the local universe. With the aid of gravitational telescopes to measure individual objects' brightness profiles to 10\\% accuracy, the study of the high-redshift size-mass relation may be extended by an order of magnitude or more beyond existing surveys at the low-mass end, thus providing a new observational test of galaxy formation models. ", "introduction": "\\label{sect:intro} Galaxies do not appear in arbitrary combinations of luminosity, mass and shape, but instead obey empirical scaling relations (such as the Fundamental Plane for early-type galaxies). Explaining the origin, and cosmic evolution, of the scaling relations is a fundamental goal of galaxy formation theories. As far as disk galaxies are concerned, the hierarchical structure formation scenario predicts a correlation between size and stellar mass, with width depending on the distribution of the initial spin of the dark halos \\citep{F+E80}. At any given mass, the expected distribution of sizes is well-approximated by a log-normal distribution. Qualitatively, this prediction is quite robust, although the exact forms of the correlation and the distribution depend on the details of baryonic processes such as energy feedback from star formation and bulge instability \\citep{MMW98,She++03,Ton++06,Dut++07,S+B07}. Therefore, measuring the shape and width of the correlation provides not only a test of the standard paradigm, but also valuable information on the poorly-understood baryonic processes happening at sub-galactic scales. From an empirical point of view, the relation between size, luminosity (or equivalently surface brightness) and stellar mass is well established for disk galaxies in the local Universe \\citep[\\eg][]{She++03,Dri++05}. Analysis of suitable objects in the Sloan Digital Sky Survey shows that at any given mass (luminosity) the distribution of galaxies is indeed well-approximated as log-normal, although the scaling with mass of the characteristic size and the width of the distribution are non-trivial. Defining disk galaxies as those being well-fit by a single \\sersic component with index $n<2.5$, \\citet{She++03} find that above a characteristic stellar mass ($\\log M_{*,0}/\\msun\\sim 10.6$ corresponding to approximately M$_{r,0}=-20.5$), size scales rapidly with stellar mass ($R\\sim M_*^{0.39}$) and the scatter is relatively small ($\\sigma_{\\ln R}\\sim 0.34$). Below the characteristic stellar mass the correlation flattens ($R\\sim M_*^{0.14}$) and the scatter increases significantly ($\\sigma_{\\ln R}\\sim 0.47$). At intermediate redshift ($0.1\\lsim z \\lsim 1$) the nature and interpretation of the size-luminosity or size-mass relation is more uncertain. Several authors \\citep[\\eg][]{Fer++04,Bar++05,Tru++06,Mel++06} have used Hubble Space Telescope images to determine the sizes of intermediate and high ($z \\gsim 1$) redshift galaxies, down to the resolution and completeness limits of HST (roughly equivalent to 1~kpc and 10$^{10} \\msun$). Recent studies conclude, taking selection effects into account, that there is significant evolution in the size-luminosity relation \\citep{Bar++05,Tru++06,Mel++06}. However, it is hard to disentangle luminosity evolution from size evolution, to ensure that samples at different redshifts are directly comparable, and to compare results from different studies, as the selection criteria are often similar but not identical (\\eg color vs. morphology; morphology determined via \\sersic index vs.\\ bulge to disk decomposition vs.\\ concentration parameter vs. visual classification). Overall, it appears that disk galaxy evolution cannot be explained by pure luminosity or pure size evolution, but requires a combination of both. In contrast, the relation between size and stellar mass appears to have changed very little since $z\\sim1$ \\citep{Bar++05}, much less than would be expected in the naive model where stellar mass and size are proportional to the virial mass and radius (and hence size is expected to scale as $H(z)^{-\\frac{2}{3}}$, where $H(z)$ is the Hubble parameter). Rather, galaxies appear to be growing ``inside-out'' in scale radius as their stellar mass increases such that the size-mass relation is preserved over cosmic time \\citep{Bar++05}. It has been suggested that galaxy evolution models that take into account the ever-increasing concentration of dark matter halos, and the further effect of baryons via adiabatic contraction could provide the physics required to reproduce the observed trend \\citep{Som++06}, although this may make it more difficult to reproduce simultaneously other scaling laws, for example the Tully-Fisher \\citep{T+F77} relation \\citep{Dut++07}. Lower mass ($M_* \\lsim 10^{10} \\msun$) galaxies are even less well understood. While the local size-mass relations of \\citet{She++03} for low ($n<2.5$) and high ($n>2.5$) \\sersic index objects diverge, the interpretation of \\sersic index as a morphological galaxy classifier becomes more uncertain at lower masses \\citep[e.g.\\ ][]{CCD92,TBB04}. At the same time, the measurement of the structural parameters themselves becomes harder as the galaxy size decreases. Nevertheless, such small galaxies are important objects to understand: the luminous compact blue galaxies first noted by \\citet{K+K88} appear in large numbers at intermediate redshifts in deep HST images \\citep[\\eg][]{Noe++06,Raw++07}, but evolve very rapidly to vanishing abundance in the local universe. What becomes of these objects, which represent sites of small-scale but vigorous star formation, is a topic of some debate, with dwarf spheroids \\citep[\\eg][]{Koo++94,Noe++06} and the bulges of disk galaxies \\citep[\\eg][]{Ham++01,Raw++07} the principle candidates. Gravitational lensing is a powerful tool with which to extend the investigation of scaling laws over cosmic time \\citep[\\eg][]{Tre07}. On the one hand, the lensing geometry provides a precise and almost model-independent measure of total mass of the lens galaxy. Since the lens galaxies are mostly early-type galaxies (or the bulges of spirals), this gives a new handle on the mass profile of these systems \\citep{T+K04,Koo++06} and hence, for example, on the relationship between stellar and total mass \\citep{Bol++07}. On the other hand, the background source is typically magnified by a factor of $\\sim$10, mostly in the form of a stretch along the azimuthal direction. While lensing preserves surface brightness, the increase in apparent size of the lensed source means that the number of pixels at any one surface brightness also increases, such that the isophotes are observed at higher signal-to-noise. Thus, gravitational lenses act as natural telescopes, allowing one to gain a factor of $\\sim10$ in sensitivity and spatial resolution, and thus improve markedly our ability to study the size and dynamical mass (through rotation curves) of intermediate and high redshift galaxies. For example, studies of the internal structure of faint blue galaxies~\\citep{Ell97}, and in particular the most compact of these \\citep{Koo++94}, are currently limited by the resolution of HST \\citep{Phi++97}. When magnified by a gravitational lens, such objects become well-resolved. Thanks to the dedicated efforts of several groups, the number of known gravitational lenses is increasing dramatically: it is now possible to envision statistical studies of relatively large sample of lensing or lensed galaxies in the near future. In this paper we present multi-color high-resolution images of the gravitational lens system \\lensname \\citep{Bol++06}, obtained with both the Hubble Space Telescope and with the Laser Guide Star Adaptive Optics (LGSAO) System on Keck II. The scientific goal of the analysis of this case study is two-fold. First, we perform a detailed comparison of the results of the lens modeling across bands, showing that -- when a bright nearby star is available for tip-tilt correction and conditions are favorable -- the most important parameters can be measured with comparable accuracy with HST and Keck-LGSAO. Second, we exploit this particular cosmic telescope to achieve super-resolution of the source galaxy. \\citep[See][ for Keck LGSAO observations of a lens with a point-like source.]{McK++07} With a lens magnification of $\\mu \\gsim 10$, the resolution of the HST and Keck images ($\\sim 0\\farcs1$ FWHM) corresponds to a physical scale of ($0.66{\\rm kpc} / \\mu \\approx 0.05{\\rm kpc}$) at the redshift of the source {\\zs~=~\\zsvalue}, comparable to the resolution attainable from the ground when studying galaxies in the Virgo Cluster in $1$~arcsec seeing. We derive the \\sersic index, size, and stellar mass of the source, and show that using gravitational telescopes the size-mass relation may be extended by an order of magnitude in size with respect to current studies, thus allowing one to probe, for example, whether the change in slope and intrinsic scatter below the characteristic mass persists to higher redshifts. This paper is organized as follows. After describing the observations in section~\\ref{sect:obs}, we outline in sections~\\ref{sect:psf} and~\\ref{sect:subtraction} two sources of systematic error and our strategies for dealing with them, before explaining our modeling methodology in section~\\ref{sect:modelmethod}. In sections~\\ref{sect:modelresults} and ~\\ref{sect:source} we present our results, which are then discussed (section~\\ref{sect:discuss}) before we draw conclusions in section~\\ref{sect:conclude}. Throughout this paper magnitudes are given in the AB system. We assume a concordance cosmology with matter and dark energy density $\\Omega_m=0.3$, $\\Omega_{\\Lambda}=0.7$, and Hubble constant H$_0$=70 kms$^{-1}$Mpc$^{-1}$. ", "conclusions": "\\label{sect:conclude} We find that high quality images from \\nirc are capable of providing very similar precision on simple lens and source model parameters to typical datasets from \\acs and \\nicmos. The data themselves contain information about the most appropriate PSF model to use, to the extent that a set of nearby unsaturated stars can be fruitfully compared using suitable statistics that are sensitive to the goodness-of-fit. We estimate that even for the LGSAO imaging this way of modeling the PSF allows a photometric precision of 0.05 mag. However, the calibration of isothermal However, the calibration of \\emph{isothermal} galaxy-scale gravitational lenses as cosmic telescopes is very likely limited by the subtraction of the lens galaxy light. We estimate that this procedure introduces up to 0.1 magnitudes of systematic error into the source galaxy photometry. However, this is still smaller than the error introduced by the assumption of an isothermal density profile for the lens itself. With this in mind we draw the following conclusions about the source behind \\lensname: \\begin{itemize} \\item Our photometry is robust enough to permit a reconstruction of the SED, and we find a stellar mass of (\\stellarmassvalue). This is a factor of 5 smaller than the completeness limit of the GEMS disk galaxy analysis of \\citet{Bar++05}, and also smaller than the least massive spheroid at this redshift studied by~\\citep{McI++05}. \\item The \\sersic profile parameters of the source can be measured to high accuracy. We find an effective radius of (\\revalue) ($\\approx 0.09$ arcsec with $\\sim10$\\% accuracy), and a \\sersic index of (\\bluesindexvalue) in the \\Iband ($\\sim$ rest-frame~B), and that these values change little over the rest-frame optical range. \\item This very small galaxy lies approximately 3-sigma below the local size-mass relation for disks. However, it shares the properties of the smallest of the compact narrow emission line galaxies of \\citet{Koo++94}, and, despite its low \\sersic index, is more typical of the dwarf early-type galaxies observed in the Virgo cluster~\\citep{Fer++06} and the ``elliptical'' galaxies studied by~\\citet{McI++05} at high redshift. \\end{itemize} While the planned statistical analysis of a large sample of lensed galaxies will rely on the detailed understanding of the selection function, it is clear that the magnifying effect of gravitational lenses allows us to extend current size-mass studies to smaller sizes and lower masses than would otherwise be available, posing fresh challenges to models of galaxy formation and evolution." }, "0710/0710.2318_arXiv.txt": { "abstract": "% Proposed mechanisms for the formation of km-sized solid planetesimals face long-standing difficulties. Robust sticking mechanisms that would produce planetesimals by coagulation alone remain elusive. The gravitational collapse of smaller solids into planetesimals is opposed by stirring from turbulent gas. This proceeding describes recent works showing that ``particle feedback,\" the back-reaction of drag forces on the gas in protoplanetary disks, promotes particle clumping as seeds for gravitational collapse. The idealized streaming instability demonstrates the basic ability of feedback to generate particle overdensities. More detailed numerical simulations show that the particle overdensities produced in turbulent flows trigger gravitational collapse to planetesimals. We discuss surprising aspects of this work, including the large (super-Ceres) mass of the collapsing bound cluster, and the finding that MHD turbulence aids gravitational collapse. ", "introduction": "Coagulation is the dominant mechanism for the growth of dust grains via van der Waals forces \\citep{dt97} and for the growth of solid protoplanets by gravitational binding \\citep{gls04}. But sticking is difficult in the ~mm--km size range, as confirmed by an extensive body of experimental work \\citep{wb06}. Moreover, incremental growth of planetesimals leads to the rapid ($\\sim10^2$ yr) inspiral of particles near a meter in size. An alternative hypothesis \\citep[hereafter GW]{saf69, gw73} proposes that km-sized planetesimals formed from the gravitational collapse of smaller solids. This mechanism overcomes the sticking and radial drift obstacles in one fell swoop. However stirring by turbulent gas opposes gravitational collapse. GW noted that their dense particle midplane would trigger a turbulent boundary layer via Kelvin-Helmholtz instabilities. \\citet{sw80} argued that this particle-driven turbulence (an example of drag force feedback) would stir up the particle midplane enough to prevent gravitational collapse. This appeared to be a fatal flaw of the GI hypothesis, since particle settling was a self-limiting process. \\citet{sek98} and \\citet{ys02} showed that particles could actually help their cause of becoming planetesimals. When the surface density of solids relative to gas is larger (by factors of few) than solar abundances, then vertical shear is no longer strong enough to overcome the anti-buoyancy of the dense midplane layer. The ability of particles to stir themselves is limited. However these works did not model the back reaction of drag forces on the gas in detail. Indeed most analyses of midplane Kelvin-Helmholtz instabilities \\citep[with the exception of][]{jhk06} reduce the particle and gas dynamics to a simplified set of equations which omit relative motion and prevent a detailed examination of the effects of drag forces. Section \\ref{sec:SI} describes the streaming instability \\citep[hereafter YG]{yg05} which takes the opposite approach of neglecting stratification and vertical shear to consider two-way drag forces in a self-consistent, if incomplete, model. This idealized system shows that particle feedback triggers spontaneous particle clumping. The detailed 3D simulations of \\citet[hereafter JOMKHY]{nature07} include vertical stratification, self-gravity, MHD turbulence and multiple particle sizes, see \\S\\ref{sec:KS} These models show that clumps of 15 -- 60 cm boulders, augmented by feedback effects, collapse gravitationally into bound clusters, which should continue to contract into planetesimals. ", "conclusions": "The long-standing mystery of planetesimal formation suddenly appears much less daunting, if no less fascinating. For decades, the obstacles seemed insurmountable: low sticking efficiencies, rapid radial migration and the inevitability of turbulent stirring. Now it is clear that turbulent stirring does not necessarily prevent gravitational collapse of solid particles into planetesimals. As discussed in \\S\\ref{sec:surprise}, more turbulence can promote gravitational collapse in some cases! The breakthoughs in particle-gas dynamics are related to the role of particle feedback. Once thought only to hinder particle settling by triggering Kelvin-Helmholz instabilities, we now realize that feedback triggers and augments particle clumping. The link between feedback and clumping has many pillars of support: the idealized streaming instability (YG, YJ, JY), Kelvin-Helmholz instabilities with uniform rotation \\citep{jhk06}, and now 3D stratified models of Keplerian disk midplanes (JOMKHY and its supplement). Much work and many uncertainties remain. The large initial sizes assumed by JOMKHY might be unrealistic, either because coagulation stalls at smaller sizes, or if less violent gravitational collapse occurs first (see work on dissipative gravitational collapse by \\citealp{war76,y05a}). The fact that most primitive undifferentiated meteorites betray no structures larger than mm-sized chondrules suggest this concern may be valid. However, particles which are small and tightly coupled to the gas disk are less effected by the dynamical clumping mechanisms discussed here, at least at the resolutions studied to date. Perhaps more surprises are needed." }, "0710/0710.1631_arXiv.txt": { "abstract": "We present a comprehensive study of accretion activity in the most underdense environments in the universe, the voids, based on the SDSS DR2 data. Based on investigations of multiple void regions, we show that Active Galactic Nuclei (AGN) are definitely common in voids, but that their occurrence rate and properties differ from those in walls. AGN are more common in voids than in walls, but only among moderately luminous and massive galaxies ($M_r < -20$, log $M_*/M_{\\sun} < 10.5$), and this enhancement is more pronounced for the relatively weak accreting systems (i.e., $L_{\\rm [O III]} < 10^{39}$ erg s$^{-1}$). Void AGN hosted by moderately massive and luminous galaxies are accreting at equal or lower rates than their wall counterparts, show lower levels of obscuration than in walls, and similarly aged stellar populations. The very few void AGN in massive bright hosts accrete more strongly, are more obscured, and are associated with younger stellar emission than wall AGN. These trends suggest that the accretion strength is connected to the availability of fuel supply, and that accretion and star-formation co-evolve and rely on the same source of fuel. Nearest neighbor statistics indicate that the weak accretion activity (LINER-like) usually detected in massive systems is not influenced by the local environment. However, H {\\sc ii}s, Seyferts, and Transition objects are preferentially found among more grouped small scale structures, indicating that their activity is influenced by the rate at which galaxies interact with each other. These trends support a potential H{\\sc~ii}$\\rightarrow$Seyfert/Transition Object$\\rightarrow$LINER evolutionary sequence that we show is apparent in many properties of actively line-emitting galaxies, in both voids and walls. The subtle differences between void and wall AGN might be explained by a longer, less disturbed duty cycle of these systems in voids. ", "introduction": "The regions that are apparently devoid of galaxies \\citep{kir81} and clusters \\citep{ein80}, the voids, are arguably the best probes of the effect of the environment and cosmology on galaxy formation and evolution. If, as suggested by the standard cosmological paradigm, structure in the present-day universe formed through hierarchical clustering, with small structures merging to form progressively larger ones, galaxies in the currently most underdense regions must be the least ``evolved'' ones, as they must have formed at later times than those in the dense regions. Therefore, void and cluster galaxies must follow different evolutionary paths. Disturbing processes like stripping and harassment, that operate preferentially in crowded environments, should occur rarely in voids. Studies of the properties of the void galaxies, in contrast to those in relatively crowded regions, or walls, should provide some of the strongest constraints for distinguishing the intrinsic properties, which characterize a galaxy when it is first assembled, from properties that have been externally induced, over the whole history the universe: the ``nature versus nurture'' problem. Statistically significant conclusions regarding the distinctness of the void galaxies relative to those in denser regions, the ``walls'' hereafter, emerged only recently, with the advent of large surveys such as SDSS and 2dF. Such data, and in particular SDSS, offered for the first time the possibility to find and analyse both photometrically and spectroscopically, large samples of extremely low density regions (i.e., $\\delta\\rho/\\rho < -0.6$ measured on a scale of $7 h^{-1}$ Mpc, Rojas et al. 2004, 2005), and allowed for accurate estimates of the void galaxy luminosity and mass functions \\citep{hoy05, gol05}. These studies show that void galaxies are fainter, bluer, have surface brightness profiles more similar to those of late-type systems, and that their specific star formation rates are higher than those in denser regions. The mass and luminosity functions are found to be clearly shifted towards lower characteristic mass and fainter magnitudes ($M^*$). Moreover, the faint end slopes of the wall and void luminosity functions are very similar which suggests that voids are not dominated by an excess population of low-luminosity galaxies. Consistently, no significant excess in the amount of dark matter is apparent. This means that, although largely devoid of light, the most underdense regions conform to a galaxy formation picture which is clearly not strongly biased. All these peculiarities demonstrate that the cosmological evolution of void systems is different from that of those living in environments of average cosmic densities. Given the tight correlations between the mass of black holes ($M_{\\rm BH}$) and the dispersions and the masses of the galactic bulges within which they reside \\citep{mag98, fer00, geb00, mar03}, one would then expect that the growth of massive BHs in galaxy centers (and therefore the accretion process within active galactic nuclei, AGN), also differs among distinct environs. Extension of environmental studies of AGN properties to extreme regions like cosmic voids is thus crucial to understand the co-evolution of galaxies and their central BHs. Moreover, while there is general agreement that the growth of black holes must be closely related to galaxy assembly \\citep{silk98, kau00, beg05}, there is no consensus as to how exactly accretion and star formation are coupled. The void galaxies could be, arguably, the best test-bed for understanding whether these processes are synchronized, or precede one another, and whether feedback from the actively growing BHs facilitates star formation (e.g., by dynamically compressing gas clouds through radio jets), or suppresses it (e.g., by blowing away the gas). To date, studies of the spectral properties of the void AGN remain limited to individual voids, e.g., the Bootes void \\citep{kir81}, permitting the identification of only a few AGN among only a few dozen void galaxies \\citep{cru02}. Quite surprisingly, such investigations find that the AGN fraction and their emission-line properties are similar in voids and in their field counterparts. Moreover, their associated stellar populations appear to share similar characteristics in the two extreme environs. The conclusions of these studies are based on small number statistics and do not however exclude the hypothesis that the void emission-line activity, whether originating in star-formation or accretion, could be connected with, e.g., filaments within voids; such structures would provide local environs similar to those in the field. The present SDSS samples of voids and void galaxies offer us for the first time the possibility to test and observationally constrain such ideas. It is important to note that previous investigations of the environmental dependence of nuclear activity in the relatively nearby universe do not reach the extreme spatial densities representative of voids. For example, in \\citet{kau04}, the lowest density regions include over 25\\% of the galaxies, which is more than 3 times more galaxies than the true void regions encompass. Their conclusions are interesting, and an extension of such an investigation at truly low densities is clearly desirable. In particular, it is important to quantify the degree to which the finding that, at fixed stellar mass, twice as many galaxies host strong-lined AGN in low-density regions than in high, extends to cosmic voids. Our work provides such an analysis. We employ in this work the most accurately classified samples of voids identified within SDSS to date, that yield $\\sim 10^3$ void galaxies. Motivated by our recent study on the AGN clustering phenomena \\citep{con06a}, which shows that there are differences in the large scale structure of active galaxies, and that their clustering amplitude correlates with their strength or rate of accretion, and possibly with the availability of fuel, we compare void and wall active galaxies of different types as classified based on their emission-line properties. Through such a comparison we aim to understand: 1) how the large and small-scale structures influence accretion onto their central black holes, and 2) to what degree AGN activity is triggered by interactions or mergers between galaxies. To answer these questions, we investigate the occurrence rate of different types of spectrally defined AGN, and how their accretion activity relates to their associated black hole mass, the mass and the age of their associated stellar populations, host morphology, brightness, and nearest-neighbor distance. We organize the paper as follows. In Section ~\\ref{data} we present the void and wall sample selection, and the spectral classification we use in defining various types of actively emitting galaxies. We compare and discuss the AGN occurrence rate in voids and walls, both globally and at fixed host properties in Section ~\\ref{fractions}. We examine in Section ~\\ref{voidaccretion} the accretion rates, the fuel supply, and the properties of the associated star-formation in void and wall actively line emitting systems, while in Section ~\\ref{nnprop} we discuss potential differences in their small scale environments. We summarize our findings in ~\\ref{discussion}, and discuss the possible implications on the nature of the power sources in the low luminosity AGN, the AGN--host connection, and current models of galaxy formation. In particular, we show empirical evidence for a possible evolutionary sequence that links different types of strong line-emitting galaxies defined based on their spectral characteristics. Throughout this work, unless otherwise noted, we assume $\\Omega_m = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and $H_0 = 100h$ km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "\\label{discussion} \\subsection{\\sc Summary of Results} \\label{results} We use the largest sample of voids and void galaxies yet defined to investigate the galactic nuclear accretion phenomenon in the most underdense regions, in relation to their more populous counterparts. By employing spectroscopic and photometric data based on the SDSS DR2 catalog, and in particular measurements available in the Garching catalog, we conduct a comparative analysis of void and wall systems of different radiative signature. We find that all types of Low Luminosity AGN exist in void regions. However, their occurrence rate and intrinsic properties show variation from their wall counterparts. The differences between the wall and void AGN seem to be driven by the properties of their hosts, which are correlated with (or governed by) their small scale environment. Following is a summary of our main results: (i) Among moderately bright or fainter galaxies ($10 < log (M_*/M_{\\sun}) < 10.5$, $M_r \\ga -20$), the rate of occurrence of AGN is higher in voids than in walls. The most common accretion activity in voids is of medium power, with $L_{\\rm [O III]} \\sim 10^{38}$ erg s$^{-1}$. For the more luminous massive hosts, due to small number statistics, the relative prevalence of accretion activity in voids versus walls remains poorly constrained. (ii) The majority of void AGN, which are hosted by $M_r \\ga -20$ galaxies, have the tendency to accrete at lower rates than in those in walls. This behavior seems related to the fact that the void systems show less obscuration and, perhaps, less dense emitting gas. That the stellar populations associated with void and wall AGN are similarly aged suggests that fuel might be equally available for accretion in void and wall galaxies of similar properties (i.e., $M_r$), but that fuel is less efficiently driven towards the nucleus in void galaxies. (iii) The few void AGN hosted by bright, massive galaxies ($M_r \\la -20$, log $M_*/M_{\\sun} > 10.5$) are LINERs that show peculiarly higher accretion rates, larger amounts of obscuring matter and more recent star-formation than in their wall counterparts. These particular systems reinforce the general trends other objects show: higher accretion rates are invariably associated with younger stellar populations and higher obscuration. These trends suggest that the amount of obscuration could be a measure of the available fuel for both star formation and accretion. (iv) The radio activity of line-emitting galaxies appears both less frequent and weaker in voids than in walls. Were we able to support these differences with statistically significant measurements, they would imply that central radio activity in wall systems, including H {\\sc ii}s, is more pronounced because it builds on contributions from accretion that remains optically obscured and therefore undetected. (v) Nearest neighbor statistics show that the type of emission-line activity is correlated with the small-scale local environment. The star-forming regions (the H{\\sc~ii}s), populate the most crowded sub-regions of voids while populating relatively sparse regions in walls; both are environments where low-mass galaxies recently formed. The weakly active galaxies (LINERs) live within the clusters in walls but the most rarefied regions in voids. This finding is puzzling and suggests that these systems, which are generally old, were probably not aware of their environments when they formed. Actively accreting systems (Seyferts and possibly the Ts) inhabit intermediate regions, which are relatively dense galaxy neighborhoods in voids but are of average density in walls. (vi) These correlations among the type and strength of galactic nuclear activity, incidence rates of different types, and their small and large scale environments, suggest an H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER evolutionary scenario in which interaction is responsible for propelling gas towards the galaxy centers, triggering star formation and feeding the active galactic nucleus. \\subsection{\\sc The H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER Evolutionary Sequence} \\label{sequence} Figure ~\\ref{h2stl} illustrates how various intrinsic and host properties of actively line-emitting galaxies follow this H {\\sc ii}$\\rightarrow$S/T$\\rightarrow$LINER sequence. The early stages of such objects manifest themselves as H{\\sc~ii}, as the accreting source remains heavily embedded in dust. As the star-burst fades in time, the dominance of the Seyfert-like excitation in systems of generally small but actively accreting black holes becomes more evident. Successive evolution reveals aging stellar populations associated with objects spectrally classified as Transition objects, that are still showing signs of accretion, followed by LINERs, whose stars are predominantly old and whose accretion onto already grown-up BHs is close to minimal. Note that this H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER progression is very similar in walls and voids. The lower accretion rates and the higher frequency of actively accreting systems in void versus wall galaxies of similar properties indicate a potential delay in the AGN dominance phase within voids. Thus, void AGN progress through the H {\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER sequence more slowly, while the sequence is similar for both void and wall galaxies. This picture fits well the observed properties of each type of galaxy nucleus: \\begin{itemize} \\item That H{\\sc~ii}-type of activity is significantly more frequent in voids than in walls suggests that their void-like environments, in which they have closer (both 1st and 3rd) neighbors than other types of objects have, are essential in triggering their activity. In other words, close encounters that produce either harassment, or major and/or minor mergers may be an important cause for igniting both accretion and star formation. \\item Seyferts' environments in both voids and walls are intermediate between those of H{\\sc~ii}s and LINERs, regardless of their brightness. If the Seyfert-like activity is triggered by interactions, probably the same ones that turn on the H{\\sc~ii}s, there must be a time lag between the onset of the star-burst and when accretion becomes dominant, or simply observable. Such a time interval corresponds to a period of aging of the stellar population (as seen in the differences in $D4000$ and H$\\delta_{\\rm A}$ between H{\\sc~ii}s and Seyferts), when the post starburst fuel becomes increasingly available for accretion. Moreover, this progression develops relatively uninterrupted in voids, and therefore, possibly, at a slower pace than it would in walls where close encounters or other types of interactions can either accelerate or terminate it. \\item Void and wall Ts are barely distinguishable in their physical characteristics. In both voids and walls, their nn-statistics and intrinsic and host properties are intermediate between those of LINERs and H{\\sc~ii}s, for any given range of $M_r$. Their BHs are apparently growing (their accretion activity is stronger than that of Ls but weaker than that of Seyferts), their fuel supply seems plentiful (they are found in some of the most obscured systems), and they are associated with (quite massive) stellar populations that are generally younger than those of LINERs, but older than the majority of star-forming systems. It is among Ts however that the most massive void accreting BHs are observed; this might suggest that, in the proposed H {\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER sequence, massive void galaxies reach the low accretion rate (i.e., LINER) phase later than in walls. \\item Whether Seyferts or Transition objects are first in this sequence remains ambiguous. Both their intrinsic properties and nn-statistics are very similar, and remain intermediate between those of H {\\sc~ii}s and LINERs. The H$\\alpha$/H$\\beta$ Balmer decrements and the $d_{\\rm 1nn}$ are the only parameters that show a ``jump'' in the otherwise smooth H{\\sc~ii}$\\rightarrow$S$\\rightarrow$T$\\rightarrow$LINER sequence manifested by other properties of these systems. Further investigations of these objects should address the differences between them in terms of a possible evolutionary progression. \\item Although we do not provide any quantitative estimates of the time spent in or during the various phases we propose here, this this analysis shows that both void and wall galaxies follow the same cycle. The AGN evolution does not affect the gravitational environment. To the contrary, it is the environment that sets the time scale for evolution along such a sequence; it seems to take longer to march through the different phases in voids than in walls, but the physics is the same. The large scale clustering is consistent with this picture: LINERs are {\\it now} more clustered because objects in dense regions underwent the H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER evolution more quickly; the higher rate of galaxy-galaxy interactions speeds up the way AGN proceed through the sequence. Hosts whose central regions are now in the H {\\sc~ii} phase will always be less clustered than current LINERs. \\end{itemize} Although far from being complete, this proposed evolutionary sequence is engaging and offers a comprehensive picture for the co-evolution of AGN and their host galaxies. The broad idea that mergers trigger star formation and that the AGN appears afterwards, in fact shutting off the star formation because of feedback, has been discussed previously in the literature. For example, N-body simulations by \\citet{byr86} and \\citet{her95} showed that interactions drive gas toward the nucleus and can produce intense star formation followed by an AGN. More recent state-of-the-art hydrodynamical models \\citep{dimat05, spr05, hop05, hop06} show that during mergers, the BH accretion peaks considerably {\\it after} the merger started, and {\\it after} the star-formation rate has peaked. However, whether early bright quasars and later, dimmer AGN obey similar physics needs still to be addressed. The H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$L sequence that this study reveals, based on the smooth alignment of several of their spectral properties, may be the first empirical evidence for an analogous duty cycle in high redshift bright systems and in nearby galaxies hosting weak quasar-like activity. This scenario can also accommodate the rather inconclusive findings regarding the role of mergers in activating AGN: their hosts do not show evidence for bars \\citep{mul97, lai02} or disturbances caused by galaxy-galaxy interactions \\citep{mal98}, exhibit morphologies very similar to those of field galaxies \\citep{derob98a, derob98b}, and pair counting in both optical and IR remain inconclusive as possible excess of companions are sometimes found \\citep{dah84, raf95}, but not always \\citep{fue88, lau95}. Moreover, \\citet{sch01, sch04} show that claims of evidence of interactions in the literature could be attributed to selection effects. The H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER cycle suggests that the majority of AGN might be detected only a certain period after the interaction, allowing time for the starburst to fade and for the BH accretion to gain strength. One might argue that that other forms of evolution could also exist for these emission-line galaxies, as opposed to this simple progression. We note that this proposed sequence does not imply that every HII galaxy at the present epoch must necessarily become a LINER; it is certainly possible that some systems go through H{\\sc~ii}-only phases, or L-only phases, which might not be part of the larger progression. We would however like to emphasize that the timescales necessary to transform from one galaxy type to the next are quite reasonable. At a first glance, this is somewhat surprising given the relatively large range in BH masses ($2 \\times 10^7 M_{\\sun}$ for H{\\sc~ii}s, to $2 \\times 10^8 M_{\\sun}$ for LINERs, inferred from $\\sigma_*$'s assuming, e.g., Tremaine et al. 2002), and their significantly low accretion rates ($L/L_{\\rm Edd} \\la 0.005$; see Figure ~\\ref{h2stl}, and note that log $L[{\\rm O III}]/\\sigma_*^4 = -1$ corresponds to approximately $L/L_{\\rm Edd} = 0.05$, according to Kewley et al. 2006); apparently, for a canonical value for the accretion efficiency, i.e., 10\\%, it takes approximately a Hubble time to e-fold in BH mass. The key is in the fact that the low luminosity AGN accrete inefficiently, as very little energy generated by accretion is radiated away (the optically thin cooling time of the gas is longer than the inflow time). Studies show that, in these cases, the radiative efficiency can be as low as $10^{-6}$ \\citep{ree82, nar94, qua03}. With a (not a necessarily extreme) efficiency value of $10^{-5}$ then, the e-folding time in the BH mass is $\\approx 4.5 \\times 10^2/l$ years, and thus, as little as few Myrs for, e.g., $l = L/L_{\\rm Edd} = 0.0005$. This (not necessarily extreme) example shows then that such an evolutionary scenario is actually physically possible. The peculiarity of LINER environments is a new and intriguing result and clearly shows that this type of activity is not controlled by its surrounding environment. Galaxies hosting LINERs could be associated with high initial density peaks in the dark matter distribution, which evolved subject to their large scale environment. Being generally massive systems to start with, LINERs' hosts would be prone to accreting material around them. In voids, this ``cleaning'' enterprise would contribute to emptying the already rarefied neighboring space, leaving little or insufficient material for future formation of (massive, bright) galaxies; they would thus end up in the most underdense void neighborhoods. In walls, and in particular within clusters, accretion of surrounding material would make a small difference as the matter density is higher. Simulations of dark matter halos and correlations with the properties of their inhabiting galaxies should be able to address these ideas. Further tests of this H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER evolutionary scenario are clearly needed. These tests require larger samples that allow separation into different morphological subsamples, and observables that parametrize the galaxy morphology better than the concentration index. Moreover, we need better constraints on the BH masses and consequently the Eddington rates for the H{\\sc~ii}s in particular, or the late type galaxies in general. When available, analysis of such parameters would shed light on the assumed coevality of star-formation and Seyfert-like BH accretion in centers of galaxies, removing ambiguities regarding the initial stages of the H{\\sc~ii}$\\rightarrow$S/T$\\rightarrow$LINER progression. \\vspace{1cm} SPECIAL NOTE: During the review process of this paper, we became aware of another piece of work which introduces the same idea of an evolutionary sequence, ``from star formation via nuclear activity to quiescence`` \\citep{sch07}. Interestingly, although their general approach, analysis, and samples used, are quite different from ours, the measurements on which the evidence for a ``star-forming, transition object, Seyfert, LINER'' sequence is built shares a common set of parameters with those used in our analysis: stellar velocity dispersions, ages of starbursts, and reddening. The global frame in which this evolutionary sequence in presented is however definitely different. While Schawinski et al. take a step forward toward understanding this possible time sequence among early-type galaxies and provide an in-depth analysis of the possible time scales involved in this picture, our paper offers a broader perspective on how such an evolutionary sequence constrains the galaxy evolution models as we provide important links to the environment. \\vspace{1cm}" }, "0710/0710.3634_arXiv.txt": { "abstract": "It is believed that quark matter can exist in neutron star interior if the baryon density is high enough. When there is a large isospin density, quark matter could be in a pion condensed phase. We compute neutrino emission from direct Urca processes in such a phase, particularly in the inhomogeneous Larkin-Ovchinnikov-Fulde-Ferrell (LOFF) states. The neutrino emissivity and specific heat are obtained, from which the cooling rate is estimated. ", "introduction": "\\label{introduction} The phase structure of quantum chromodynamics (QCD) is one of the most challenging problem in particle and nuclear physics. We schematically illustrate in Fig.\\ \\ref{phase} the phase diagrams in $T-\\m_B$ and $T-\\m_I$ plots, where $T$, $\\m_B$ and $\\m_I$ are the temperature, baryon and isospin (or equivalently electron) chemical potentials respectively. In the $T-\\m_B$ diagram, the left panel of Fig.\\ \\ref{phase}, the hadronic phase locates at low $T$ and low $\\m_B$ region and undergoes a phase transition or a crossover to the deconfined quark phase at certain critical temperature $T_c$ or baryon chemical potential $\\m_{Bc}$ of the orders of $T_c\\sim 200$ MeV or $\\m_{Bc}\\sim 1$ GeV. At very high temperature, the quark-gluon-plasma (QGP), made of free quarks and gluons, forms. At asymptotically high $\\m_B$ but low $T$, the ground state of QCD is the color-flavor-locked (CFL) superconductor \\cite{alford1999} where the condensation of quark pairs spontaneously breaks color and chiral symmetries. At intermediate $T$ and $\\m_B$, although quarks and gluons are deconfined they are still strongly coupled. In this regime, many QCD phases are proposed in recent years, such as, at low $T$, two-flavor color superconductivity (2SC) \\cite{Ruester:2004eg}, gapless 2SC (g2SC) \\cite{shovkovy2003}, gapless CFL (gCFL)\\cite{alford2004}, spin-1 color superconductor \\cite{iwasaki1995,schafer2000,schmitt2002}, kaon condensation in the CFL phase \\cite{kaon-condensation}, et al.. For reviews of color superconductivity, see, e.g. Ref. \\cite{csc-reviews}. There may exist resonance states at intermediate $T$, such as strongly coupled QGP (sQGP) \\cite{shuryak2006} at low $\\m_B$ or the pseudo-gap phase at moderate $\\m_B$ \\cite{kitazawa2005}. In the $T-\\m_I$ diagram, the right panel of Fig.\\ \\ref{phase}, the hadron phase is in the region with low $T$ and low $\\m_I$ while the QGP phase locates at very high $T$. At low $T$, when $\\m_I$ grows above the value of the pion mass $m_\\p$, the ground state turns out to be a Bose-Einstein condensation (BEC) of pions, as $\\m_I$ increases further and is larger than about 230 MeV, the pion BEC crossover smoothly into the BCS superfluid of quark-anti-quark pairs with the condensate $\\langle \\ub i\\g_5 d\\rangle$ or $\\langle \\db i\\g_5u\\rangle$ \\cite{son2001}. At intermediate $T$, resonance states such as sQGP may occur at low $\\m_I$ while the states with strong fluctuations of thermally excited mesons or Cooper pairs are possible at moderate $\\m_I$. \\begin{figure}[!htb] \\begin{center} \\includegraphics[width=6cm]{phase1.eps} \\includegraphics[width=6cm]{phase2.eps} \\caption{The schematic phase diagrams of QCD on $T-\\m_B$ and $T-\\m_I$ planes.} \\label{phase} \\end{center} \\end{figure} In this paper we consider the quark matter cores in neutron stars. The neutrino emission from direct Urca processes $d\\ra u+e^-+\\nb, u+e^-\\ra d+\\n$ is the most efficient way of cooling in quark matter. Sch\\\"afer and Schwenzer summarized the neutrino emissivities and specific heats for a variety of color superconducting phases of quark matter \\cite{schafer2004}. Due to beta equilibrium the isospin chemical potential $\\m _I$ is nonzero, quark matter could be a pion BEC ($\\m_I<230$ MeV) or a BCS superfluid ($\\m_I > 230$ MeV) when $\\m _I > m_\\p$ \\cite{son2001}. On the other hand, the baryon chemical potential is still large (in the order of 1 GeV) and makes a big mismatch between the fermi surfaces of the pairing quarks $u(\\bar u)$ and $\\bar d(d)$, thus the BEC or BCS state is gapless. Such a gapless phase is stable in the BEC region but unstable in the BCS one with respect to the formation of nonzero LOFF momentum. In a previous work \\cite{huang2007}, we have studied the neutrino emissivity and cooling rate due to Urca processes for the gapless pion condensed quark matter in the BEC region. In this paper we study the neutrino emission in the LOFF phase. We work in two-flavor case in the moderate baryon density, where the role of strange quarks is not important. Our units are $\\hbar=k_B=c=1$ except particular specifications. As a convention, we denote a 4-momentum as $K^\\mu =(k_0,\\mathbf{k})$, and its 3-momentum magnitude as $k=|\\mathbf{k}|$. \\section {Quark Propagator\\label{quark}} Our starting point is the two flavor Nambu-Jona-Lasinio Lagrangian of QCD \\begin{eqnarray} \\label{NJL} \\cl=\\jb(i\\g^\\m\\pt_\\m+\\m\\g_0-m_0)\\j+g[(\\jb\\j)^2+(\\jb i\\vec{\\t}\\g_5\\j)^2], \\end{eqnarray} where $\\j=(u,d)^{\\rm T}$ is the quark fields, $g$ is the coupling constant and $\\vec{\\t}$ is the Pauli matrices. We have introduced the chemical potential matrix in flavor space, $\\m=\\mathrm{diag}(\\m_u, \\m_d)=(\\m+\\d\\m, \\m-\\d\\m)=(\\m_B/3+\\m_I/2, \\m_B/3-\\m_I/2)$ with $\\m_B, \\m_I$ the baryon and isospin chemical potential respectively. We assume that the $\\b$-equilibrium is reached $\\m_d=\\m_u+\\m_e$, which gives $\\m_I=-\\m_e$. In the chiral limit and without the chemical potentials, the Lagrangian (\\ref{NJL}) respects the symmetry $U_B(1)\\otimes SU_V(2)\\otimes SU_A(2)$ corresponding to the baryon number, isospin vector and pseudovector conservation respectively. However, the presence of the chemical potentials explicitly break the isospin symmetry down to $U_V(1)$ with the conserved quantum number of $\\t_3$, and chiral symmetry down to $U_A(1)$ with the conserved quantum number of $i\\g_5\\t_3$. By introducing the chiral condensate $\\s=-2g\\lan\\jb\\j\\ran$ and pion condensates $\\p^-=\\D e^{-2i\\bl\\cdot\\bx}=4g\\langle\\ub i\\g_5 d\\ran$, $\\pi^+=\\D e^{2i\\bl\\cdot\\bx}=4g\\langle\\db i\\g_5 u\\ran$, we arrive at the mean field Lagrangian \\begin{eqnarray} \\cl_\\mf=\\jb\\lb\\begin{array}{cc}i\\g^\\m \\pt_\\m+\\m_u\\g_0-m & \\p^+i\\g_5 \\\\ \\p^-i\\g_5 & i\\g^\\m \\pt_\\m+\\m_d\\g_0-m \\end{array}\\rb\\j -\\frac{\\s^2+\\D^2}{4g}, \\end{eqnarray} with the effective quark mass $m=m_0+\\s$. There are mismatches in Fermi surfaces of anti-u and d quarks or anti-d and u quarks by the baryon chemical potential, hence pion condensates with nonzero total momentum or the LOFF states may be favored. The formation of the condensate $\\s\\sim\\lan\\jb\\j\\ran$ breaks the $U_A(1)$ chiral symmetry spontaneously with the Goldstone boson $\\p_0\\sim\\jb i\\g_5\\t_3\\j$, and that of pion condensates $\\D\\sim \\langle\\db i\\g_5u\\ran\\sim \\langle\\ub i\\g_5d\\ran$ break the $U_V(1)$ isospin symmetry spontaneously. The translational and rotational symmetries are spontaneously broken by nonzero LOFF momentum $\\bl$. The partition function of the system can be written as a functional integral \\begin{equation} Z=\\int [d\\jb ] [d\\j ] \\exp{\\left( \\int_0^\\b d\\t\\int d^3\\bx\\cl_\\mf\\right) } . \\end{equation} Rewriting the quark fields via a gauge transformation which leaves the partition function unchanged, $\\c_u(x)=u(x)e^{-i\\bl\\cdot\\bx}$, $\\c_d(x)= d(x)e^{i\\bl\\cdot\\bx}$ (we still call $\\c _{u,d}$ quark fields), the inverse quark propagator in flavor and momentum space reads \\begin{eqnarray} \\label{s-inverse} S^{-1}(K)=\\lb\\begin{array}{cc}\\g^\\m K_\\m-\\bl\\cdot\\g+\\m_u\\g_0-m & \\D i\\g_5 \\\\ \\D i\\g_5 & \\g^\\m K_\\m+\\bl\\cdot\\g+\\m_d\\g_0-m \\end{array}\\rb. \\end{eqnarray} Note that $K^\\m$ represents the 4-momentum of the $\\c$ fields instead of the $\\j$ fields. The propagator is written as \\begin{eqnarray} \\label{propagator} S(K)=\\lb\\begin{array}{cc} S_{uu}(K) & S_{ud}(K) \\\\ S_{du}(K) & S_{dd}(K) \\end{array}\\rb. \\end{eqnarray} A straightforward calculation from Eq. (\\ref{s-inverse}) gives the four elements \\begin{eqnarray} S_{uu}(K)&=&\\frac{(\\g^\\m K_{+\\m}+m)(K_-^2-m^2-\\D^2)+2\\D^2\\g^\\m l_\\m}{(K_+^2-m^2+\\D^2)(K_-^2-m^2+\\D^2)-\\D^2[(K_++K_-)^2-4m^2]},\\non S_{dd}(K)&=&\\frac{(\\g^\\m K_{-\\m}+m)(K_+^2-m^2-\\D^2)-2\\D^2\\g^\\m l_\\m}{(K_+^2-m^2+\\D^2)(K_-^2-m^2+\\D^2)-\\D^2[(K_++K_-)^2-4m^2]},\\non S_{ud}(K)&=&\\frac{(\\g^\\m K_{+\\m}+m)(K_-^2-m^2-\\D^2)+2\\D^2\\g^\\m l_\\m}{(K_+^2-m^2+\\D^2)(K_-^2-m^2+\\D^2)-\\D^2[(K_++K_-)^2-4m^2]}\\non &&\\times \\frac{\\g^\\m K_{-\\m}-m}{K_-^2-m^2}i\\g^5\\D,\\non S_{du}(K)&=&\\frac{(\\g^\\m K_{-\\m}+m)(K_+^2-m^2-\\D^2)-2\\D^2\\g^\\m l_\\m}{(K_+^2-m^2+\\D^2)(K_-^2-m^2+\\D^2)-\\D^2[(K_++K_-)^2-4m^2]}\\non &&\\times \\frac{\\g^\\m K_{+\\m}-m}{K_+^2-m^2}i\\g^5\\D,\\non \\end{eqnarray} where $K_\\pm^\\m=(k_0+\\m\\pm\\d\\m,\\bk\\pm\\bl)$ and $l^\\m=(\\d\\m, \\bl)$. The excitation spectra of the quasi-particles can be obtained by solving the equation $\\det S^{-1}(k_0,\\bk)=0$ or equivalently the roots of the denominator of the propagator for $k_0$, \\begin{eqnarray} 0&=&(K_+^2-m^2+\\D^2)(K_-^2-m^2+\\D^2)-\\D^2[(K_++K_-)^2-4m^2]\\non &\\approx&[(k_0+\\m+\\ve^-_{\\bk,\\bl})^2-(\\ve_{\\bk,\\bl}^+ +\\d\\m)^2 -\\D^2]\\non &&\\times [(k_0+\\m-\\ve_{\\bk,\\bl}^-)^2-(\\ve_{\\bk,\\bl}^+-\\d\\m)^2-\\D^2], \\end{eqnarray} where $\\ve_{\\bk,\\bl}^\\pm=(E_{\\bk+\\bl}\\pm E_{\\bk-\\bl})/2$ with $E_\\bk\\equiv\\sqrt{\\bk^2+m^2}$. To arrive at the last line, we have taken the assumption that both $\\D$ and $\\bl$ are small comparing to the quark Fermi momenta in the LOFF phase. We have four excitation branches, $E_r^a(\\bk,\\bl)=-r\\sqrt{(\\ve_{\\bk,\\bl}^+ - a\\d\\m)^2+\\D^2}-(\\m - a\\ve_{\\bk,\\bl}^-)$, with $a,r=\\pm$. Taking the same approximation to the numerators in the elements of the quark propagator, we neglect the terms proportional to $\\D^2\\bl$. Thus we rewrite the elements of the quark propagator as, \\begin{eqnarray} \\label{suu-sdd} S_{uu}(K)&\\simeq&\\sum_{a,r=\\pm}\\frac{B_{r}^{a}(\\bk ,\\bl )\\L^{a}_{\\bk+\\bl }\\g_0}{k_0-E_r^a(\\bk,\\bl)},\\non S_{dd}(K)&\\simeq&\\sum_{a,r=\\pm}\\frac{B_{-r}^{a}(\\bk ,\\bl )\\L^{-a}_{\\bk -\\bl }\\g_0}{k_0-E_r^a(\\bk,\\bl)}, \\end{eqnarray} where we have introduced the energy projectors, \\begin{equation} \\L^a_{\\bk }=\\frac{1}{2}\\left[1+a\\frac{\\g_0(\\g\\cdot\\bk+m)}{E_\\bk}\\right], \\end{equation} and the Bogoliubov coefficients, \\begin{equation} \\label{bog-coeff} B^{a}_r(\\bk,\\bl)=\\frac 12 \\left[1-ar\\frac{\\varepsilon ^+_{\\bk,\\bl}-a\\delta\\mu} {\\sqrt{(\\varepsilon ^+_{\\bk,\\bl}-a\\delta\\mu)^2+\\Delta^2 }}\\right], \\end{equation} ", "conclusions": "" }, "0710/0710.3896_arXiv.txt": { "abstract": "We analyze the mean rest-frame ultraviolet (UV) spectrum of Type Ia Supernovae (SNe) and its dispersion using high signal-to-noise Keck-I/LRIS-B spectroscopy for a sample of 36 events at intermediate redshift ($\\overline{z}$=0.5) discovered by the Canada-France-Hawaii Telescope Supernova Legacy Survey (SNLS). We introduce a new method for removing host galaxy contamination in our spectra, exploiting the comprehensive photometric coverage of the SNLS SNe and their host galaxies, thereby providing the first quantitative view of the UV spectral properties of a large sample of distant SNe~Ia. Although the mean SN~Ia spectrum has not evolved significantly over the past 40\\% of cosmic history, precise evolutionary constraints are limited by the absence of a comparable sample of high quality local spectra. The mean UV spectrum of our $z\\simeq$0.5 SNe~Ia and its dispersion is tabulated for use in future applications. Within the high-redshift sample, we discover significant UV spectral variations and exclude dust extinction as the primary cause by examining trends with the optical SN color. Although progenitor metallicity may drive some of these trends, the variations we see are much larger than predicted in recent models and do not follow expected patterns. An interesting new result is a variation seen in the wavelength of selected UV features with phase. We also demonstrate systematic differences in the SN~Ia spectral features with SN lightcurve width in both the UV and the optical. We show that these intrinsic variations could represent a statistical limitation in the future use of high-redshift SNe~Ia for precision cosmology. We conclude that further detailed studies are needed, both locally and at moderate redshift where the rest-frame UV can be studied precisely, in order that future missions can confidently be planned to fully exploit SNe~Ia as cosmological probes. ", "introduction": "\\label{sec:introduction} Supernovae of Type Ia (SNe~Ia) are now well-established as cosmological distance indicators. In addition to the original surveys by the Supernova Cosmology Project \\citep[SCP;][]{1997ApJ...483..565P,1999ApJ...517..565P} and the High-Z Supernova Search Team \\citep{1998ApJ...507...46S,1998AJ....116.1009R}, a new generation of SN~Ia surveys is underway both locally \\citep{2002SPIE.4836...61A,2005coex.conf..525L,2006PASP..118....2H} and at higher redshifts \\citep{2006A&A...447...31A,2007ApJ...659...98R,2007ApJ...666..694W}. Despite the availability of independent probes of the presence and properties of dark energy from studies of the cosmic microwave background \\citep{2007ApJS..170..377S} and galaxy redshift surveys \\citep{2002MNRAS.330L..29E,2005MNRAS.362..505C,2005ApJ...633..560E}, the luminosity distance--redshift relation for SNe~Ia provides the only {\\it direct} evidence for a cosmic acceleration. The detection and characterization of dark energy, via measurements of the average cosmic equation of state parameter $<$$w$$>$, requires the precision measurement of SNe~Ia to redshifts $z\\simeq$0.5--1, sampling the epoch of cosmic acceleration \\citep{2006A&A...447...31A}. However, more precise constraints on the nature of dark energy, for example evidence for any variation in $w$ with redshift, requires extending these studies to redshift $z$$>$1 \\citep{2004ApJ...607..665R,2007ApJ...659...98R} where the early effects of deceleration may be detectable. As projects are developed which plan to probe SNe~Ia beyond $z$=1 for this purpose \\citep[e.g][]{2005NewAR..49..346A,2006SPIE.6265E..67B}, it becomes important to understand the possible limitations of using SNe~Ia as distance probes. Key issues relating to the diversity of SNe~Ia and their possible evolution with redshift as a population, together with the limiting effects of dust and/or color corrections to their photometric properties, are particularly crucial to understand. Several local studies \\citep{1995AJ....109....1H,1999AJ....117..707R, 2000AJ....120.1479H,2001ApJ...554L.193H, 2005A&A...433..807M,2005ApJ...634..210G} have already indicated correlations between SN~Ia properties and host galaxy morphologies. More recently, \\citet{2006ApJ...648..868S} have shown that the properties of distant SNe~Ia appear to be a direct function of their local stellar population, with the distribution of light curve widths and hence peak luminosities correlating with the host galaxy specific star-formation rate. This work also determined that the rate of SNe~Ia per unit stellar mass of their host galaxies is larger in actively star-forming galaxies, suggesting many must be produced quite rapidly in recently formed stellar populations, perhaps suggestive of more than one progenitor mechanism. The authors conclude that SNe~Ia may well be a bimodal or a more complex population of events \\citep[see also][]{2005ApJ...629L..85S,2006MNRAS.370..773M}. Such diversity in the properties of SNe~Ia could have far-reaching implications, particularly if the {\\it mix} of mechanisms or delay times within the broad population gradually changes with look-back time \\citep[e.g.][]{2006MNRAS.370..773M,2006ApJ...648..868S,2007astro.ph..1912H}. These recent developments, which illustrate how improved precision reveals new physical correlations in the SN~Ia population, raise the broader question of whether future SN~Ia experiments might be limited in precision by variations of a systematic nature within the population, for example with redshift, which cannot be removed via empirical correlations. Detailed local surveys such as the LOSS/KAIT \\citep{2005coex.conf..525L} and CfA surveys \\citep{1999AJ....117..707R,2006AJ....131..527J} have presented valuable data on the homogeneity and trends in the SNe~Ia population. Further promising work is being undertaken via the Supernova Factory \\citep{2002SPIE.4836...61A} and the Carnegie Supernova Project \\citep{2006PASP..118....2H}. Important though these continued programs will be, they are insufficient to address all possible concerns about the use of SNe~Ia as precision tools in cosmology. Comparable studies at intermediate redshift\\footnote{Defined here to represent the range 0.2$<$$z$$<$0.7.} will be particularly important in order to address questions relating to possible evolutionary effects and environmental dependencies. In addition, it is not always practical at low redshift to cover the full wavelength range necessary to test for systematic trends. In this paper we analyze high signal-to-noise ratio rest-frame ultraviolet (UV) spectra of a large sample of intermediate redshift SNe~Ia drawn from the Canada-France-Hawaii Telescope Supernova Legacy Survey \\citep[SNLS;][]{2006A&A...447...31A}, a rolling search which is particularly effective for locating and studying events prior to their maximum light. Our aim is obtain a substantially higher signal-to-noise in the spectra than that typically obtained during spectroscopic programs to type SNe and measure redshifts. We target the UV wavelength region because in this wavelength region the SN spectrum is thought to provide the most sensitive probe of {\\it progenitor metallicity} \\citep[e.g.][]{1998ApJ...495..617H,2000ApJ...530..966L}, a variable which may shed light on the possibility of progenitor evolution. The time-dependent UV spectrum is also needed in estimating ``cross-band'' $k$-corrections, particularly at redshifts $z>1$ where optical bandpasses probe the rest-frame near-UV \\citep{2004ApJ...607..665R,2007ApJ...659...98R}. Little is known about the properties and homogeneity of the UV spectra of SNe~Ia, largely because of the absence of suitable instruments for studying this wavelength range in local events. Although some local SN~Ia UV spectra are available from International Ultraviolet Explorer (IUE) or Hubble Space Telescope (HST) satellite data \\citep[e.g.][]{1991ApJ...371L..23L,1993ApJ...415..589K,1995ESASP1189.....C}, the bulk of the progress now possible in this area can be provided from optical studies of intermediate-redshift events with large ground-based telescopes. The goals of this paper are thus to address the question of both the diversity and possible physical evolution in the intermediate redshift SN~Ia family. We compare the rest-frame UV behavior of local SNe~Ia with that derived for intermediate-redshift ($z\\simeq$0.5) events where the rest-frame UV enters the region of high efficiency of the Keck LRIS-B spectrograph. We also study the degree to which the UV spectra at intermediate redshift represent a homogeneous population, independent of other variables such as the physical environment and light curve stretch. A plan of the paper follows. In $\\S$~\\ref{sec:selection-cfhtls-sne} we introduce the salient features of the SNLS and our method for selecting SNe~Ia for detailed study. In $\\S$~\\ref{sec:keck-observations} we discuss the Keck spectroscopic observations and their reduction, including the treatment of host galaxy subtraction and flux calibration. In $\\S$~\\ref{sec:analyses} we consider our sample with respect to the broader set of SNe found by SNLS, ensuring it is a representative subset in terms of various observables, and discuss existing local UV spectra. In $\\S$~\\ref{sec:results}, we undertake the detailed analysis. First we compare the UV spectra found in our sample with those found locally. We then examine the diversity of intermediate-redshift SNe~Ia in various ways and correlate the UV variations with the light curves of the SNe and the properties of the host galaxies. We discuss these trends in terms of progenitor mechanisms in $\\S$~\\ref{sec:discussion} and examine the implications in terms of possible long term limitations of SNe~Ia as probes of dark energy. We also present the mean phase-dependent SN Ia spectrum and its uncertainties for use in future work. ", "conclusions": "\\label{sec: conclusions} We summarize our findings as follows: \\begin{enumerate} \\item{} We have secured high signal-to-noise ratio Keck spectra for a sample of 36 intermediate redshift SNe~Ia, observed at various phases, spanning the redshift range 0.15$ 6.5\\times 10^{-8}$ using just the TT, TE and EE power spectra. ", "introduction": "The inflationary paradigm is strongly supported by observations of the cosmic microwave background (CMB) made by the COBE and WMAP satellites~\\cite{Smoot:1992td,Spergel:2006hy}. However, it is still a paradigm in search of a specific model based on fundamental physics. Brane inflation~\\cite{Dvali:1998pa} has emerged as one of the most popular ways of embedding inflation within string theory. It uses a natural candidate for the inflaton: the field which describes the brane-antibrane separation. This field has a non-trivial potential due to the attractive brane-antibrane interaction which is flattened by the effect of the compactification of the extra dimensions \\cite{Burgess:2001fx}, and the geometry of the branes \\cite{GarciaBellido:2001ky}. Cosmic strings are also a natural occurrence within this model and are result of inhomogeneities in the tachyon field of the brane-antibrane pair \\cite{C_Strings1,C_Strings2}. These are represented by a complex field with a non-trivial potential which supports the formation of codimension 2 defects which have been shown to exhibit the correct properties of lower dimension branes \\cite{Tachyon_Strings}. Since reheating of the universe in these model proceeds via the annihilation of the brane-antibrane pair, the formation of cosmic strings is expected at the end of the inflation and the strings will be naturally located at the bottom of the throat as their tension is also warp-factor dependent. The most complete model of brane inflation incorporating moduli stabilization has been proposed in ref.~\\cite{Kachru:2003sx}. In this model inflation happens naturally as a mobile brane falls down the warped throat, being attracted by an antibrane stuck at the bottom of the throat. Antibranes have a warp-factor dependent potential, and therefore minimize their energy by moving to the bottom of the throat, the region of strongest warping. The mobile brane is not affected by the warping. One can understand this by considering the warped geometry as being generated by a large stack of branes. An antibrane is attracted to the stack and will move towards it, that is towards largest warping. The brane feels no force from the stack of branes (it is BPS with respect to them), only from the antibrane located at the bottom of the throat. This situation can be described by a simple scalar field, since in this model all moduli are stabilized and only the brane-antibrane separation is evolving. Constraints on the cosmic string tension, $G\\mu$, come from a variety of observations. Of most interest here are those which are most robust. In particular, we will concentrate on the constraints which result from their inclusion as a sub-dominant component in the angular power spectrum of anisotropies of the cosmic microwave background (CMB)~\\cite{CHMb,WB,Wyman:2005tu,Seljak:2006bg,Bevis:2006mj,Battye:2006pk,Bevis:2007gh}. As pointed out recently~\\cite{Battye:2006pk,Bevis:2007gh} in the context of the third year WMAP data, larger values of the spectral index of density fluctuations, $n_{\\rm s}$, are compatible with observations if a sub-dominant string component with around $5-10\\%$ of the large-scale amplitude is included. This is a generic feature of any inflation model which produce strings. In ref.~\\cite{Battye:2006pk} accurate constraints on the coupling constant, $\\kappa$ and mass scale, $M$, relevant to the simplest models of supersymmetric F- and D-term hybrid inflation were computed, taking into account the fact that the observed power spectrum is described by 3 parameters, $G\\mu$, $n_{\\rm s}$ and the power spectrum amplitude, $P_{\\cal R}$, each of which can be derived from $\\kappa$ and $M$. In this model dependent approach more powerful constraints are possible. We will adapt the same approach, where relevant to the case of brane inflation in this paper. ", "conclusions": "\\label{sec:discussion} In this paper we have presented limits and constraints on the parameters (and derived parameters) of brane inflation models in the slow-roll regime when a cosmic string component is included in the fitting. Important aspects of the results are an increased range of acceptable values of $n_{\\rm s}$, limits on $\\gamma$, which is related to the volume of the internal space, $\\beta$ the inflaton mass parameter and a constraint on the value of $\\phi_{\\rm e}36$ these constraints will apply. Finally we comment that there may be limits on $G\\mu$ which come from pulsar timing if the cosmic string network achieves scaling by the creation of loops and the subsequent emission of radiation (see ref.~\\cite{Caldwell:1996en} and references therein). We caution that these should be considered to be less robust since they are more strongly effected by the small-scale dynamics, such as loop formation, of the cosmic string network. This is not completely understood in the case of standard cosmic strings in 3+1 dimensions; the situation with respect to the higher dimension cosmic strings is even less clear. Nonetheless, limits on the energy density in gravitational waves, $\\Omega_{\\rm g}h^2$, have substantially improved in recent times. Probably the most reliable limit is $\\Omega_{\\rm g}h^2<2\\times 10^{-8}$ at frequencies $f=2\\times 10^{-9}{\\rm Hz}$~\\cite{Jenet:2006sv}. Limits on $G\\mu$ from such a bound were considered in ref.~\\cite{Battye:2006pk} and they are dependent on the loop production size relative to the horizon $\\alpha$. If $\\alpha<10^{-4}$ then $G\\mu>10^{-6}$ is excluded and hence the limit from CMB anisotropy is strongest, whereas for $\\alpha>10^{-4}$ one finds that $G\\mu> 10^{-10}/\\alpha$ is excluded, which would be tighter than the CMB limit for $\\alpha>10^{-3}$. It is clear that an improved understanding of the loop production mechanism coupled with expected improvements in the bound on $\\Omega_{\\rm g}h^2$ could lead to a more powerful constraint than is expected from PLANCK." }, "0710/0710.3772_arXiv.txt": { "abstract": "Procyon A is a bright F5IV star in a binary system. Although the distance, mass and angular diameter of this star are all known with high precision, the exact evolutionary state is still unclear. Evolutionary tracks with different ages and different mass fractions of hydrogen in the core pass, within the errors, through the observed position of Procyon A in the Hertzsprung-Russell diagram. For more than 15 years several different groups have studied the solar-like oscillations in Procyon A to determine its evolutionary state. Although several studies independently detected power excess in the periodogram, there is no agreement on the actual oscillation frequencies yet. This is probably due to either insufficient high-quality data (i.e., aliasing) or due to intrinsic properties of the star (i.e., short mode lifetimes). Now a spectroscopic multi-site campaign using 10 telescopes world-wide (minimizing aliasing effects) with a total time span of nearly 4 weeks (increase the frequency resolution) is performed to identify frequencies in this star and finally determine its properties and evolutionary state. ", "introduction": "The bright F5 subgiant Procyon A is the primary of an astrometric binary system with a white dwarf in a 40 year orbit. Procyon A is the brightest northern-hemisphere asteroseismology candidate with well-determined characteristics, such as distance, mass and angular diameter. Brown {\\etal} \\cite{brown1991} were among the first to observe an excess power between 0.5 and 1.5 mHz in radial velocity observations confirmed by several other radial velocity studies, e.g. \\cite{martic1999}, \\cite{bouchy2002}, \\cite{kambe2003}, \\cite{martic2004}, \\cite{eggenberger2004}, \\cite{claudi2005}, \\cite{leccia2007}. So far these studies have independently revealed detections of power excess, but there is no agreement yet on the actual oscillation frequencies. This may be due to aliases present in the spectral window, short mode lifetimes, shifts from the asymptotic relation due to avoided crossings, or any combination of these factors. Although the frequencies are not yet known in detail, most studies obtain a large frequency spacing of $55 \\pm 1$ $\\mu$Hz. \\begin{table}[h!] \\caption{\\label{stelpar} Stellar parameters of Procyon A from \\cite{allende2002}.} \\begin{center} \\begin{tabular}{lr@{$\\pm$}l} \\br Mass [M$_{\\odot}$]& 1.42 & 0.06\\\\ T$_{\\rm eff}$ [K] & 6512 & 49 \\\\ Radius [R$_{\\odot}$] & 2.071 & 0.020\\\\ $\\rm[Fe/H]$ [dex] & $-$0.09 & 0.03 \\\\ $v_{\\rm rot}\\sin i$ [km\\,s$^{-1}$] & 3.16 & 0.50\\\\ \\br \\end{tabular} \\end{center} \\end{table} Another point of discussion is the fact that \\cite{matthews2004} did not detect any power excess in their MOST photometry and published Procyon A as a flat liner. Other photometric studies such as WIRE \\cite{bruntt2005} and a reanalysis of the MOST 2004 data \\cite{regulo2005} claim to detect power excess in the same region as the radial velocity studies. A recent reanalysis of the MOST 2004 data and the analysis of new MOST data taken in 2005 reinforce the null detection of p-modes \\cite{guenther2007}. This issue is discussed more extensively by \\cite{bedding2005}, who claim that the non-detection of oscillations in Procyon by the MOST satellite is fully consistent with the ground based radial-velocity studies due to a combination of several noise sources and the low photometric amplitude of the oscillations. Procyon A is in a very interesting evolutionary state near the end of its main sequence life. The stellar parameters (see Table~\\ref{stelpar}) are all known to high precision, but several evolutionary tracks with different ages and different hydrogen core mass fractions overlap, within the errors, with its position in the HR-diagram. The exact evolutionary state of a star can be revealed by means of asteroseismology and therefore the oscillation frequencies are needed. To determine the actual frequencies of Procyon A a ground-based multi-site campaign using 10 telescopes with a total time span of nearly 4 weeks was performed from December 28, 2006 till January 22, 2007. Here we present first results of this campaign. In Section 2 the campaign is described, while Section 3 discusses the way the data of the different telescopes are combined, and a final power spectrum is obtained. Some concluding remarks and future prospects are provided in Section 4. ", "conclusions": "The Procyon campaign presented here is the largest spectroscopic campaign, so far, aimed at the detection and identification of solar-like oscillations. Using 10 telescopes over a time span of nearly 4 weeks provided us with a unique data set with high time coverage and frequency resolution. The details of the data processing methods will be fully described by \\cite{arentoft2008}, while the oscillation frequencies extracted from the full data set acquired during the spectroscopic Procyon campaign will be presented by \\cite{bedding2008}. \\begin{figure} \\begin{center} \\includegraphics[width=\\linewidth]{powersmall.eps} \\caption{\\label{power}Final power spectrum of Procyon A with window-optimised weights. The data are high-pass filtered to remove low-frequency drifts.} \\end{center} \\end{figure}" }, "0710/0710.1777_arXiv.txt": { "abstract": "Studies of stellar magnetism at the pre-main sequence phase can provide important new insights into the detailed physics of the late stages of star formation, and into the observed properties of main sequence stars. This is especially true at intermediate stellar masses, where magnetic fields are strong and globally organised, and therefore most amenable to direct study. This talk reviews recent high-precision ESPaDOnS observations of pre-main sequence Herbig Ae-Be stars, which are yielding qualitatively new information about intermediate-mass stars: the origin and evolution of their magnetic fields, the role of magnetic fields in generating their spectroscopic activity and in mediating accretion in their late formative stages, and the factors influencing their rotational angular momentum. ", "introduction": "\\subsection{Magnetism and rotation in the main sequence A and B stars} Between about 1.5 and 10 M$_{\\odot}$, at spectral types A and B, about 5 \\% of main sequence (MS) stars have magnetic fields with characteristic strengths of about 1kG. Such stars also show important chemical peculiarities and are thus usually called the magnetic chemically peculiar Ap/Bp stars. The strength of the magnetic fields of these stars cannot be explained by an envelope dynamo as in the sun. Until now, the most reliable hypothesis has been to assume a fossil origin for these magnetic fields. This hypothesis implies that the stellar magnetic fields are relics from the field present in the parental interstellar cloud. Its also implies that magnetic fields can (at least partially) survive the violent phenomena accompanying the birth of stars, and can also remain throughout their evolution and until at least the end of the MS, without regeneration. According to the fossil field model, we should observe magnetic fields in some pre-main sequence (PMS) stars of intermediate mass, the so-called Herbig Ae/Be stars. However no magnetic field was observed up to recently in these stars \\cite[except HD~104237,][]{donati97}. Can we obtain some observational evidence of the presence of magnetic fields during the PMS phase of evolution, as predicted by the fossil field hypothesis? If some Herbig Ae/Be stars are discovered to have magnetic fields, is the fraction of magnetic to non-magnetic Herbig stars the same as the fraction for main sequence stars? Is the magnetic field in Herbig stars strong enough to explain the strength of that of Ap/Bp stars? Chemical peculiarities and magnetism are not the only characteristic properties observed in the Ap/Bp stars. Most magnetic MS stars have rotation periods (typically of a few days) that are several times longer than the rotation periods of non-magnetic MS stars (a few hours to one day). It is usually believed that magnetic braking, in particular during PMS evolution, when the star can exchange angular momentum with its massive accretion disk, is responsible for this low rotation \\cite[][]{stepien00,stepien02}. An alternative involves a rapid dissipation of the magnetic field during the early stages of PMS evolution for the fastest rotators, due to strong turbulence induced by rotational shear developed under the surface of the stars, as the convection do in the solar-type stars \\cite[see e.g.][]{lignieres96}. In this scenario, only slow rotators could retain their initial magnetic fields, and evolve as magnetic stars to the main sequence. So the question to be addressed is the following: does the magnetic field control the rotation of the star, or else does the rotation of the star control the magnetic field? We propose that this question can be answered by studying rotation and magnetic fields in Herbig Ae/Be stars. \\subsection{The Herbig Ae/Be stars} The Herbig Ae/Be stars are intermediate-mass pre-main sequence stars, and therefore the evolutionary progenitors of the MS A and B stars. They are distinguished from the classical Be stars by their IR emission and the association with nebulae, characteristics which are due to their young age. They display many observational phenomena often associated with magnetic activity. First, high ionised lines are observed in the spectra of some stars \\cite[e.g.][]{bouret97,roberge01}, and X-ray emission have been detected, coming from some Herbig stars \\cite[e.g.][]{hamaguchi05}. In active cool stars, many of these phenomena are produced in hot chromospheres or coronae. Some authors mentioned rotational modulation of resonance lines which they speculate may be due to rotation modulation of winds structured by magnetic field \\cite[][]{praderie86,catala89,catala91,catala99}. In the literature we find many clues of the presence of circumstellar disks around these stars, from spectroscopic data showing strong emission, and also from photometric data \\cite[e.g.][]{mannings97,mannings00}. Recently, using coronagraphic data and interferometric data, some authors have also found direct evidence of circumstellar disks around these stars \\cite[][]{grady99,grady00,eisner03}. A careful study of these disks shows that they have similar properties to the disk of their low mass counterpart \\cite[][]{natta01}, the T Tauri stars, whose the emission lines are explained by magnetospheric accretion models \\cite[][]{konigl91,muzerolle98,muzerolle01}. Finally \\citet{muzerolle04} have sucessfully applied their magnetospheric accretion model to Herbig stars to explain the emission lines in their spectra. For all these reasons we suspect that the Herbig stars may host large-scale magnetic fields that should be detectable with current instrumentation. However, many authors tried to detect such fields without much success \\cite[][]{catala93,catala99,donati97,hubrig04,wade07}. But in 2005, a new high-resolution spectropolarimeter, ESPaDOnS, has been installed at the canada-france-hawaii telescope (CFHT). We therefore decided to proceed to survey many Herbig stars in order to investigate rotation and magnetism in the pre-main sequence stars of intermediate mass. ", "conclusions": "We used the new spectropolarimeter ESPaDOnS installed at the CFHT to proceed in a survey of the Herbig stars, in order to investigate their rotation and magnetic field. We discovered four magnetic stars whose field topology is similar to the MS magnetic A-B stars. We also show that the magnetic intensities of these fields and the proportion of magnetic Herbig stars can explain the magnetic intensity and the proportion of magnetic fields among the MS stars, in the context of fossil field model. We therefore bring fundamental arguments in favour of this hypothesis. The four magnetic Herbig stars are slow rotators ($v\\sin i < 41$ km.s$^{-1}$) which supports that magnetic Herbig Ae/Be stars are the progenitors of the magnetic Ap/Bp stars. Among these magnetic stars two have very low $v\\sin i$ ($< 10$ km.s$^{-1}$) and are very young (age$ < 2.8$ Myr). Assuming these stars are true slow rotators, this implies that there exists a braking mechanism which acts very early during the PMS evolution of the intermediate mass stars. We could think that this braking mechanism has a magnetic origin, although among the undetected stars we also observe same stars with small $v\\sin i$ ($\\sim 15$ km.s$^{-1}$). The nature of the braking mechanism requires addition study. Finally it has been proposed that all Herbig stars should undergo magnetospheric accretion, as the spectra of all stars show similar emission. However, we calculated the minimum polar magnetic intensity of the magnetic Herbig stars and of two well-constrained undetected stars to get magnetospheric accretion, using three different models which have been developed for the T Tauri stars \\cite[][]{konigl91,cameron93,shu94}. Considering the intensity of the magnetic stars, as well as the maximum magnetic intensity of the two well-constrained undetected stars, if they host a magnetic field, our first conclusion is that magnetospheric accretion cannot occur in all the Herbig stars (a statistic study taking into account all the undetected stars is in progress in order to confirm this result). Therefore, either the models that we used are not well adapted to the Herbig stars, or the emission lines are not only produced by magnetospheric accretion. A thorough observational and theoretical study of the emission in the spectra, as well as the surroundings of the Herbig Ae/Be stars is necessary to better understand the interaction of these stars with their surroundings. \\vspace{1cm}" }, "0710/0710.4370_arXiv.txt": { "abstract": "Based on a simulation of galaxy formation in the standard cosmological model, we suggest that a consistent picture for Gamma-Ray Bursts and star formation may be found that is in broad agreement with observations: {\\it GRBs preferentially form in low metallicity environments and in galaxies substantially less luminous that $L_*$.} We find that the computed formation rate of stars with metallicity less than $0.1\\zsun$ agrees remarkably well with the rate evolution of Gamma-Ray Bursts observed by Swift from $z=0$ to $z=4$, whereas the evolution of total star formation rate is weaker by a factor of about $4$. Given this finding, we caution that any inference of star formation rate based on observed GRB rate may require a more involved exercise than a simple proportionality. ", "introduction": "The intriguing observational linkage between long duration Gamma-Ray Bursts (GRBs) and core-collapse supernovae (e.g., Stanek \\etal 2003; Hjorth \\etal 2003) suggests that the progenitors of GRBs may be very massive stars. This possible connection was predated by a proposed unified picture (Cen 1998). As such, it has been hoped that GRBs may be a good tracer of cosmic star formation (Wijers \\etal 1998; Totani 1999; Lamb \\& Reichart 2000; Blain \\& Natarajan 2000; Porciani \\& Madau 2001; Daigne \\etal 2006; Coward 2006; Le \\& Dermer 2007; Li 2007). However, recent observations indicate that typical GRBs may prefer relatively low metallicity environments ($\\sim 0.1\\zsun$) (Fynbo \\etal 2003; Le Floch \\etal 2003; Christensen \\etal 2004; Fruchter \\etal 2006; Stanek \\etal 2006) and host galaxies significantly less luminous than $L_*$ (Fruchter \\etal 1999,2006), although there is evidence that the actual spread in metallicity may be wide (Berger \\etal 2006; Prochaska 2006; Wolf \\& Podsiadlowski 2007). The aim of this {\\it Letter} is to first address the issue of consistency of GRB environment with respect to metallicity and galaxy luminosity, i.e., the galaxy luminosity-metallicity relation, in the context of detailed simulation of galaxy formation in the standard cosmological model. Then, we make predictions on the evolution of GRB rate with redshift and highlight a possible dramatic difference between overall star formation history and GRB rate history, if GRBs are not an unbiased tracer of star formation. In particular, if GRBs are predominantly produced by stars with metallicity $\\le 0.1\\zsun$, the GRB rate is expected in our model to rise obstinately from $z=0$ to $z\\sim 5$ by a factor of $\\sim 100$, when it flattens out towards higher redshift, whereas the overall star formation rate rises rapidly only from $z=0$ to $z\\sim 3$ and is roughly flat from $z\\ge 3$ until $z\\sim 7$. The evolution of GRB rate with redshift is thus expected to be stronger than that of star formation. ", "conclusions": "We utilize a simulation of galaxy formation in the standard cosmological model that has been shown to produce results consistent with extant observations of galaxy formation (e.g., Nagamine \\etal 2006) to shed light on the relation between GRB rate and star formation rate. We find that a consistent picture for Gamma-Ray Bursts and star formation that is in broad agreement with observations would emerge, {\\it if GRBs preferentially form in low metallicity environments and in galaxies substantially less luminous that $L_*$.} Because of the increase of metallicity with cosmic time, GRB rate consequently evolves more strongly with redshift than the overall star formation rate. We find that the observed evolution of GRB rate from $z=0$ to $z=4$ can be explained, {\\it if GRBs are primarily produced by massive stars with metallicity less than $0.1\\zsun$}, whereas an inclusion of stars with metallicity as high as $0.3\\zsun$ yields GRB rate evolution from $z=0$ to $z=4$ inconsistent with observations. Therefore, we reach the conclusion that GRBs may not be a good tracer of cosmic star formation, especially over a long timeline. As a result, a simple inference of star formation rate or its derived quantities such as the ionizing photon production rate at high redshifts, based on the observed GRB rate, should be done with caution and may require careful calibrations. \\smallskip" }, "0710/0710.4146_arXiv.txt": { "abstract": "Motivated by the increasing use of the Kennicutt-Schmidt (K-S) star formation law to interpret observations of high redshift galaxies, the importance of gas accretion to galaxy formation, and the recent observations of chemical abundances in galaxies at \\ztwo--3, I use simple analytical models to assess the consistency of these processes of galaxy evolution with observations and with each other. I derive the time dependence of star formation implied by the K-S law, and show that the sustained high star formation rates observed in galaxies at \\ztwo--3 require the accretion of additional gas. A model in which the gas accretion rate is approximately equal to the combined star formation and outflow rates broadly reproduces the observed trends of star formation rate with galaxy age. Using an analytical description of chemical evolution, I also show that this model, further constrained to have an outflow rate roughly equal to the star formation rate, reproduces the observed mass-metallicity relation at \\ztwo. ", "introduction": "The motivations of this paper are several. First, the empirical Kennicutt-Schmidt (K-S) law \\citep{s63,k98schmidt}, which states that the surface density of star formation is proportional to the surface density of gas, is widely used to interpret and describe star formation in galaxies, though its origins are not fully understood and it is just beginning to be tested at high redshift \\citep{btg+04,csn+07,bcd+07}. The K-S law is a valuable tool when gas masses are not directly measurable; it has been used to estimate the gas masses of both high redshift galaxies \\citep{ess+06mass} and local galaxies in the distant past \\citep{cjp+07}. The consequences of the K-S law for the evolution of star formation at high redshift are therefore worth considering in more detail, as is the consistency of these consequences with observations. Second, the fueling of star formation by gas accretion is an essential element of models of galaxy formation, but has largely been neglected by observers of galaxies at high redshifts, mostly because of the lack of evidence for inflow in the spectra of these galaxies. In contrast, the evidence for strong outflows in galaxies at \\ztwo--3 is well-known, most notably in the form of offsets between the redshifts of nebular emission lines, rest-frame UV absorption lines, and \\lya\\ emission (\\citealt{pss+01}, Steidel et al.\\ 2007, in prep). In spite of the lack of observations of inflow, however, its effects should be considered in the context of other observations, since a significant inflow rate would affect other, measurable properties. Finally, in recent years metallicity measurements of increasingly large samples of galaxies at $z>1$ have become possible (e.g.\\ \\citealt{kk00,pss+01,sga+04,sep+04,mlc+06}), including the detection of a mass-metallicity relation at $z>2$ \\citep{esp+06}. These measurements still suffer considerably from limitations on the methods that can be used and from calibration uncertainties, but nevertheless they offer a unique opportunity to place constraints on star formation histories and gas flows at high redshift, provided that the effects of inflow, outflow and star formation can be disentangled. Many recent studies have addressed this issue in some detail, including \\citet{ke99,kh05,d07}; and \\citet{fd07}. The goal of this paper is to formulate simple, analytical models including star formation according to the K-S law, gas inflows and outflows, and chemical evolution. We would like to assess whether or not these models are consistent with each other and with our current observational knowledge of high redshift galaxies, and to see if they might give a general picture of how gas flows, star formation and metal enrichment may proceed at high redshift. In \\S\\ref{sec:ks} I derive the explicit time dependence of star formation implied by the K-S law and test its consistency with observations of galaxies at \\ztwo. In \\S\\ref{sec:metals} I consider simple models of chemical evolution which incorporate both inflows and outflows of gas, and again test their consistency with observations of high redshift galaxies and with the results of the previous section. Some implications of the results are discussed in \\S\\ref{sec:discuss}. I adopt a cosmology with $H_0=70\\;{\\rm km}\\;{\\rm s}^{-1}\\;{\\rm Mpc}^{-1}$, $\\Omega_m=0.3$, and $\\Omega_{\\Lambda}=0.7$, and use the \\citet{c03} initial mass function (IMF). ", "conclusions": "\\label{sec:discuss} The above results provide a coherent picture in which strong star formation is sustained by the accretion of gas at approximately the gas processing rate, the outflow rate is roughly equal to the SFR, and metal enrichment is modulated by both outflows and inflows. This is not a new result; \\citet{fd07} reached many of the same conclusions using cosmological hydrodynamic simulations to reproduce the \\ztwo\\ mass-metallicity relation, and the idea of a system in which star formation is balanced by inflow dates to work by \\citet{l72} and \\citet{tl78}. Whatever the methods used to reach these conclusions, however, many questions remain about the mechanisms of gas flows and chemical enrichment at high redshift. The only quantity not yet tied to observations is the gas accretion rate, which the models require to be approximately equal to the gas processing rate. If the outflow rate is roughly equal to the SFR, the required accretion rate is $\\sim60$ \\msunyr\\ (assuming the average SFR of the UV-selected sample; \\citealt{ess+06stars}), and as much as several hundred \\msunyr\\ or higher for the most rapidly star forming galaxies. These values are in general agreement with theoretical expectations. For example, the predicted average gas accretion rates for galaxies in $10^{12}$ \\msun\\ haloes\\footnote{Clustering results indicate that the \\ztwo\\ UV-selected galaxies are typically associated with $\\sim10^{12}$ \\msun\\ haloes \\citep{asp+05}.} given by \\citet{kkwd05} are $\\sim50$ \\msunyr\\ at the relevant redshifts, rising to several hundred \\msunyr\\ for the $10^{13}$ \\msun\\ haloes expected to host the most massive galaxies (though more recent simulations indicate rates a factor of $\\sim2$ or more lower; D. Kere{\\v s}, private communication). Gas accretion and cooling rates from the semi-analytic models of \\citet{csw+06} are also of the right order of magnitude. The question remains as to why such high accretion rates have not yet been observed. One difficulty is that the velocity range of inflowing gas is likely to be much narrower than the several hundred \\kms\\ spread observed in the outflows. Another suggestion is that cold, filamentary accretion may dominate at high redshifts \\citep{kkwd05,db06}, in which case detection would depend strongly on projection effects; alternatively, the accretion could occur largely in the form of minor mergers. Another possibility is that the accreting gas may be too hot to produce signatures in the observed wavebands, or such signatures may simply be too weak to detect. Even if the hot accreting gas does produce \\ion{C}{4} emission, for example, the line would likely be weak because of the low metallicity of the gas, and it would be superposed on the already complicated \\ion{C}{4} profile. Detailed modeling of the likely line strengths would help to place limits on detectable accretion rates. Gas heated to the virial temperature of $\\sim10^6$ K must also be sufficiently cooled in order to fuel star formation; work by \\citet{yssw02} and \\citet{csw+06}, among others, discusses the mechanisms by which this might proceed. The strong star formation and gas accretion discussed herein will not continue indefinitely. Observations suggest that most of the star-forming galaxies currently detected at \\ztwo--3 will become largely passively evolving by $z\\sim1$ \\citep{asp+05,pmd+06}. Because the high observed star formation rates require accretion of new gas to sustain, a decline in the SFR is a natural consequence of the drop in accretion rates at lower redshifts predicted by theoretical models. Many theorists and observers have also proposed that AGN feedback may be responsible for the termination of star formation (e.g.\\ \\citealt{hhc+06} and references therein). If an additional mechanism to shut off star formation is required, this is a plausible candidate, as the ubiquity of outflows suggests that starburst-driven winds may regulate star formation but do not usually terminate it, and this work implies that strong accretion and outflows may operate simultaneously, or at least alternate in relatively quick succession. Finally, these results underscore the importance of metallicity measurements for understanding gas flows. Until the flows can be observed and quantified directly, measurements of gas phase abundances offer the best hope for constraining the outflow and inflow rates of galaxies at high redshift. There are still considerable difficulties associated with the measurements of these metallicities, but we look forward to improved constraints from new IR spectra and photoionization modeling, and to more direct estimates of outflow rates from detailed spectra (e.g.\\ \\citealt{psa+00}). Such measurements will give a far more robust picture of star formation, gas flows and metallicity at high redshift than these simple models can provide." }, "0710/0710.3416_arXiv.txt": { "abstract": "The \\wire\\ satellite was launched in March 1999 and was the first space mission to do asteroseismology from space on a large number of stars. \\wire\\ has produced very high-precision photometry of a few hundred bright stars ($V<6$) with temporal coverage of several weeks, including K~giants, solar-like stars, \\dss\\ stars, and $\\beta$~Cepheids. In the current work we will describe the status of science done on seven detached eclipsing binary systems. Our results emphasize some of the challenges and exciting results expected from coming satellite missions like COROT and Kepler. Unfortunately, on 23 October 2006, communication with \\wire\\ failed after almost eight years in space. Because of this sad news we will give a brief history of \\wire\\ at the end of this paper. ", "introduction": "The failure of the main mission of the \\wire\\ satellite shifted focus to the star tracker, which has a small aperture of $52$\\,mm and is equipped with a 512$\\times$512 pixel SITe CCD. During observations the main target is positioned near the middle of the CCD and the four brightest stars in the 8$\\times$8 degree field of view are also monitored. These four secondary stars are chosen automatically by the on-board software. Each observation comprises a time stamp and an 8$\\times$8 pixel window centred on the target. Two images per second are collected for each star, resulting in typically one million CCD windows in three weeks. Due to pointing restrictions two fields are observed during each \\wire\\ orbit. The duty cycle for one star per orbit is optimally 40\\%, but can be as low as 20\\%. The orbital period has decreased from 96 to 93 minutes over the cause of the mission. The filter reponse of the star tracker is not well defined but Johnson $V+R$ has been suggested \\citep{buzasi00}. For more details about \\wire\\ see Sect.~\\ref{sec:epitaph}. \\begin{figure*}[h] \\centering \\includegraphics[width=14cm]{wire_vancouver_hr_markbin.ps} \\caption{\\label{fig:hr} Hertzsprung-Russell diagram of around 100 stars observed with \\wire. The locations of the seven dEBs and the Sun are marked.} \\end{figure*} ", "conclusions": "We have presented on-going work on the observations and modelling of several eclipsing binary systems observed with the \\wire\\ satellite. The week-long temporal coverage of the targets has made it possible to study dEB systems that are notoriously difficult to observe from the ground, due to their periods being close to an integer number of days. The high precision of the photometry from \\wire\\ is a huge improvement compared to even the best photometric observations from the ground. The two important advantages of space based photometry are: \\begin{itemize} \\item No atmospheric scintillation noise \\item High stability of the \\wire\\ star tracker over periods of several weeks. \\end{itemize} These data have made it possible to push the limits on the constraints we can put on the theoretical models of these stars, covering the HR diagram from early B to late F type main sequence stars. However, at this level of precision the oscillations of the component stars, although amplitudes are small, need to be taken into account as part of the light curve analysis. In a broader context, the results presented here give an idea of the potential of secondary science on dEBs that can be done with data from the future satellite photometry missions. Many new systems will be discovered with COROT (see \\citep{michel05}) and Kepler as discussed by \\cite{koch07}. Although the new systems will be much fainter, follow-up spectroscopy and multi-band photometry can still be done from the ground." }, "0710/0710.0553_arXiv.txt": { "abstract": "The DAMA Collaboration has recently analyzed its data of the extensive WIMP direct search (DAMA/NaI) which detected an annual modulation, by taking into account the channelling effect which occurs when an ion traverses a detector with a crystalline structure. Among possible implications, this Collaboration has considered the case of a coherent WIMP-nucleus interaction and then derived the form of the annual modulation region in the plane of the WIMP-nucleon cross section versus the WIMP mass, using a specific modelling for the channelling effect. In the present paper we first show that light neutralinos fit the annual modulation region also when channelling is taken into account. To discuss the connection with indirect signals consisting in galactic antimatter, in our analysis we pick up a specific galactic model, the cored isothermal-sphere. In this scheme we determine the sets of supersymmetric models selected by the annual modulation regions and then prove that these sets are compatible with the available data on galactic antiprotons. We comment on implications when other galactic distribution functions are employed. Finally, we show that future measurements on galactic antiprotons and antideuterons will be able to shed further light on the populations of light neutralinos singled out by the annual modulation data. ", "introduction": "\\label{sec:intro} \\begin{figure}[t] \\centering \\vspace{-20pt} \\includegraphics[width=1.0\\columnwidth]{A1_global.ps} \\vspace{-30pt} \\caption{WIMP--nucleon scattering cross-section as a function of the WIMP mass. The solid (dashed) line denotes the annual modulation region derived by the DAMA Collaboration with (without) the inclusion of the channeling effect. The two regions contain points where the likelihood- function values differ more than 4$\\sigma$ from the null hypothesis (absence of modulation). These regions are obtained by varying the WIMP galactic distribution function (DF) over the set considered in Ref. \\cite{bcfs} and by taking into account other uncertainties of different origins \\cite{damalast}. The scatter plot represents supersymmetric configurations calculated with the supersymmetric model summarized in the Appendix. The (red) crosses denote configurations with a neutralino relic abundance which matches the WMAP cold dark matter amount ($0.092 \\leq \\Omega_{\\chi} h^2 \\leq 0.124$), while the (blue) dots refer to configurations where the neutralino is subdominant ($\\Omega_{\\chi} h^2 < 0.092$).} \\label{fig:00} \\end{figure} In a recent paper \\cite{damalast} the DAMA Collaboration has analyzed the data of its extensive WIMP direct search (DAMA/NaI) \\cite{dama} which measured an annual modulation effect at 6.3 $\\sigma$ C. L., by taking into account the channelling effect. This effect occurs when an ion traverses a detector with a crystalline structure \\cite{drob}. In Ref. \\cite{damalast} implications of channelling have been discussed in terms of a specific modelling of this effect for the case of the DAMA NaI(Tl) detector; it is shown that the occurrence of channelling makes the response of this detector to WIMP-nucleus interactions more sensitive than in the case in which channelling is not included. Therefore, when applied to a WIMP with a coherent interaction with nuclei, the inclusion of the channelling effect implies that the annual modulation region, in the plane of the WIMP-nucleon cross section versus the WIMP mass, is considerably modified as compared to the one derived without including channelling. The extent of the modification depends on the specific model--dependent procedure employed in the evaluation of the channeling effect \\cite{damalast}. These properties are shown in Fig. \\ref{fig:00}, where the quantity $\\sigma^{\\rm nucleon}_{\\rm scalar}$ denotes the WIMP-nucleon scalar cross-section, $\\xi = \\rho_{\\rm WIMP}/\\rho_0$ is the WIMP local fractional matter density and $m_{\\chi}$ is the WIMP mass. The dashed line denotes the annual modulation region derived by the DAMA Collaboration without including the channeling effect \\cite{dama}. The solid line shows the annual modulation region derived by the same Collaboration when the channeling effect is included as explained in Ref. \\cite{damalast}. The regions displayed in Fig. \\ref{fig:00} are derived by varying the WIMP galactic distribution function (DF) over the set considered in Ref.\\cite{bcfs} and by taking into account other uncertainties of different origins \\cite{damalast,nota1}. Fig. \\ref{fig:00} shows that the effect of taking channelling into account is that the annual modulation region modifies its contour with an extension towards lighter WIMP masses. Most remarkably, for WIMP masses $\\lsim $ 30 GeV, the WIMP-nucleon cross section involved in the annual modulation effect decreases sizeably, up to more than an order of magnitude. As mentioned before, the specific shape of the annual modulation region depends on the way in which channelling is modelled \\cite{damalast}. These features are of great importance for a specific dark matter candidate, the light neutralino, which was extensively investigated in Refs. \\cite{lowneu,lowdir,ind}. Actually, in these papers we analyzed light neutralinos, {\\it i.e.} neutralinos with a mass $m_{\\chi} \\lsim 50$ GeV, which arise naturally in supersymmetric models where gaugino mass parameters are not related by a GUT--scale unification condition. In Refs. \\cite{lowneu,lowdir} it is proved that, when R-parity conservation is assumed, these neutralinos are of great relevance for the DAMA/NaI annual modulation effect. In these papers it is also shown that in MSSM without gaugino mass unification the lower limit of the neutralino mass is $m_{\\chi} \\gsim$ 7 GeV \\cite{hp}. In Fig. \\ref{fig:00}, superimposed to the annual modulation regions is the scatter plot of the supersymmetric configurations of our model, whose features are summarized in the Appendix. One sees that, also when the channeling effect is taken into account, the light neutralinos of our supersymmetric model fit quite well the annual modulation region. In the present paper we consider the phenomenological consequences for light neutralinos when the annual modulation region is the one indicated by the solid line in Fig.\\ref{fig:00}. More specifically we examine the properties of our supersymmetric population of light relic neutralinos in terms of the possible antimatter components generated by their pair annihilation in the galactic halo. To do this, we have to resort to a specific form for the WIMP DF. We take as our representative DF a standard cored isothermal-sphere model, though we do not mean to associate to this model prominent physical motivations as compared to other forms of DFs. Analyses similar to the one we present here for the cored isothermal-sphere can be developed for other galactic models. We will comment about some of them, selected among those considered in Ref. \\cite{bcfs} (we will follow the denominations of this Reference to classify our DFs). The scheme of the present paper is the following. In Sect. II, we show how the model presented in Refs. \\cite{lowneu,lowdir,ind} fits the DAMA/NaI annual modulation regions of Ref. \\cite{damalast} for the case of the cored isothermal-sphere model. In Sect. III we combine these results with constraints derivable from available data on cosmic antiprotons; we also discuss the sensitivity of upcoming measurements on cosmic antiprotons for investigating the neutralino populations selected by the annual modulation regions. Complementary investigations by measurements of galactic antideuterons are presented in Sect. IV. Conclusions are drawn in Sect.V. The main features of the supersymmetric scheme adopted here are summarized in the Appendix. ", "conclusions": "In the present paper we have considered the annual modulation regions which the DAMA Collaboration has recently determined, by including also the channelling effect which occurs when an ion traverses a detector with a crystalline structure, such the detector of the DAMA/NaI experiment. The inclusion of the channelling effect implies that the annual modulation region is considerably modified as compared to the one derived without including channelling. The extent of the modification depends on the specific model--dependent procedure employed in the evaluation of the channeling effect. In the present paper we have considered the phenomenological consequences for light neutralinos when the annual modulation region includes the channelling effect as modelled in Ref.\\cite{damalast}. We have proved that these annual modulation data are fitted by light neutralinos which arise naturally in supersymmetric models where gaugino mass parameters are not related by the a GUT--scale unification condition. The precise range of the neutralino mass which fits the annual modulation data depends on how the WIMP galactic distribution function is modelled and on a number of other assumptions, such as those mentioned in Sect. II. As an example, we have worked out in detail the case of a cored isothermal sphere DF. For this instance, the neutralino mass stays in the range $m_{\\chi} \\simeq (7 - 30)$ GeV, for values of the local rotational velocity, $v_0$, and of the local dark matter density, $\\rho_0$, in the low-medium side of their own physical ranges, {\\it i.e.} $v_0 \\simeq$ (170 - 220) km sec$^{-1}$ and $\\rho_0 \\simeq (0.2 - 0.4)$ GeV cm$^{-3}$. Similar ranges are found also in the case of a Navarro--Frenk--White profile or for an isothermal model with a non-isotropic velocity dispersion. We have then shown that the populations of light neutralinos selected by the annual modulation regions are consistent with present data on galactic antiprotons. We have also derived the intervals of the diffusion parameters which provide this agreement in correlation with the specific galactic halo model. For instance, for neutralinos with a mass of 20 GeV and a cored isothermal model with $v_0 = 170$ km s$^{-1}$ we have $0.55 \\lsim \\delta \\lsim 0.85$ and $L \\lsim 3$ kpc when $ \\rho_0 = \\rho_0^{\\rm max} = 0.42$ GeV cm$^{-3}$; instead when $ \\rho_0 = \\rho_0^{\\rm min} = 0.20$ GeV cm$^{-3}$, $L$ may go up to 15 kpc with a range of $\\delta$ which progressively shrinks to $\\delta \\sim 0.70 - 0.75$, when $L$ increases. We have also shown that future measurements of galactic antiprotons and antideuterons will offer, together with the upcoming data from DAMA/LIBRA, very interesting perspectives for further investigating the light neutralino populations selected by the annual modulation data. In case of models with a corotating halo or with triaxial spatial distributions, not investigated in the present paper, also heavier neutralinos can be involved. Finally, a word of caution should be said concerning the fact that the distribution of WIMPs in the Galaxy could deviate from the models mentioned above, mainly because of the presence of streams and/or clumpiness. In such instances, the analysis should be appropriately adapted, along the lines discussed in the present paper." }, "0710/0710.0765_arXiv.txt": { "abstract": "CCD $VRI$ photometry is presented for SN 2002hh from 14 days after the outburst till day 347. SN 2002hh appears to be normal type IIP supernova regarding both luminosity and the shape of the light curve, which is similar to SN 1999gi. ", "introduction": " ", "conclusions": "" }, "0710/0710.0003_arXiv.txt": { "abstract": "We present a study of elemental abundances in a sample of thirteen Blue Compact Dwarf (BCD) galaxies, using the $\\sim$10--37$\\mu$m high resolution spectra obtained with Spitzer/IRS. We derive the abundances of neon and sulfur for our sample using the infrared fine-structure lines probing regions which may be obscured by dust in the optical and compare our results with similar infrared studies of starburst galaxies from ISO. We find a good correlation between the neon and sulfur abundances, though sulfur is under-abundant relative to neon with respect to the solar value. A comparison of the elemental abundances (neon, sulfur) measured from the infrared data with those derived from the optical (neon, sulfur, oxygen) studies reveals a good overall agreement for sulfur, while the infrared derived neon abundances are slightly higher than the optical values. This indicates that either the metallicities of dust enshrouded regions in BCDs are similar to the optically accessible regions, or that if they are different they do not contribute substantially to the total infrared emission of the host galaxy. ", "introduction": "Blue Compact Dwarf Galaxies (BCDs) are dwarf galaxies with blue optical colors resulting from one or more intense bursts of star-formation, low luminosities (M$_B>-$18) and small sizes. The first BCD discovered was I\\,Zw\\,18 by \\citet{Zwicky66}, which had the lowest oxygen abundance observed in a galaxy \\citep{Searle72}, until the recent study of the western component of SBS0335-052 \\citep{Izotov05}. Although BCDs are defined mostly by their morphological parameters, they are globally found to have low heavy element abundances as measured from their HII regions (1/30\\,Z$_\\odot$$\\sim$1/2\\,Z$_\\odot$). The low metallicity of BCDs is suggestive of a young age since their interstellar medium is chemically unevolved. However, some BCDs do display an older stellar population and have formed a large fraction of their stars more than 1Gyr ago \\citep[see][]{Loose85,Aloisi07}. The plausible scenario that BCDs are young is intriguing within the context of Cold Dark Matter models which predict that low-mass dwarf galaxies, originating from density perturbations much less massive than those producing the larger structures, can still be forming at the current epoch. However, despite the great success in detecting galaxies at high redshift over the past few years, bona fide young galaxies still remain extremely difficult to find in the local universe \\citep{Kunth86,Kunth00,Madden06}. This is likely due to the observational bias of sampling mostly luminous more evolved galaxies at high redshifts. If some BCDs are truly young galaxies, they would provide an ideal local laboratory to understand the galaxy formation processes in the early universe. Over the past two decades, BCDs have been studied extensively in many wavelengths using ground-based and space-born instruments \\citep [for a review see][]{Kunth00}. In the FUV, the Far Ultraviolet Spectroscopic Explorer (FUSE) has been used to study the chemical abundances in the neutral gas in several BCDs \\citep{Thuan02, Aloisi03, Lebouteiller04}. Optical spectra have been obtained for a large number of BCDs and display strong narrow emission lines resulting from the intensive star-formation processes that take place in these systems \\citep{Izotov97, Izotov99b, Pustilnik05, Salzer05}. The Infrared Space Observatory (ISO) revealed unexpectedly that despite their low metallicities, BCDs, such as SBS\\,0335-052E, could still have copious emission from dust grains \\citep{Thuan99, Madden00, Madden06, Plante02}. More recently, the Spitzer Space Telescope \\citep{Werner04} has been used to observe these metal-poor dwarf systems in order to study their dust continuum properties and the polycyclic aromatic hydrocarbon (PAH) features \\citep{Houck04b, Hogg05, Engelbracht05, Rosenberg06, Wu06, OHalloran06, Hunt06, Wu07}. Finally, radio observations have also been performed for several BCDs to study their HI kinematics and distribution \\citep{Thuan04} and thermal/non-thermal continuum emission properties \\citep{Hunt05}. \\begin{deluxetable*}{lrllllcc} \\tabletypesize{\\scriptsize} \\setlength{\\tabcolsep}{0.02in} \\tablecaption{Observing Parameters of the Sample\\label{tab1}} \\tablewidth{0pc} \\tablehead{ \\colhead{Object Name} & \\colhead{RA} & \\colhead{Dec} & \\colhead{AORKEY} & \\colhead{Observation} & \\colhead{Redshift} & \\multicolumn{2}{c}{On-source Time (Sen)}\\\\ \\colhead{} & \\colhead{(J2000)} & \\colhead{(J2000)} & \\colhead{} & \\colhead{Date} & \\colhead{} & \\colhead{SH} & \\colhead{LR} \\\\ } \\startdata Haro11 & 00h36m52.5s & -33d33m19s & 9007104 & 2004-07-17 & 0.0206 & 480 & 240 \\\\ NGC1140 & 02h54m33.6s & -10d01m40s & 4830976 & 2004-01-07 & 0.0050 & 480 & 240 \\\\ SBS0335-052E & 03h37m44.0s & -05d02m40s & 11769856 & 2004-09-01 & 0.0135 & 1440 & 960 \\\\ NGC1569 & 04h30m47.0s & +64d50m59s & 9001984 & 2004-03-01 & $\\sim$ 0& 480 & 240 \\\\ IIZw40 & 05h55m42.6s & +03d23m32s & 9007616 & 2004-03-01 & 0.0026 & 480 & 240 \\\\ UGC4274 & 08h13m13.0s & +45d59m39s & 12076032 & 2004-10-23 & 0.0015 & 120 & 56 \\\\ & & & 12626688 & 2004-11-11 & & 120 & 56 \\\\ IZw18 & 09h34m02.0s & +55d14m28s & 9008640 & 2004-03-27 & 0.0025 & 480 & 240 \\\\ & & & 16205568 & 2005-12-16 & & 2880 & 1440 \\\\ VIIZw403 & 11h27m59.9s & +78d59m39s & 9005824 & 2004-12-09 & $\\sim$ 0& 480 & 240 \\\\ Mrk1450 & 11h38m35.6s & +57d52m27s & 9011712 & 2004-12-12 & 0.0032 & 480 & 240 \\\\ UM461 & 11h51m33.3s & -02d22m22s & 9006336 & 2005-01-03 & 0.0035 & 480 & 240 \\\\ & & & 16204032 & 2006-01-14 & & 1440 & \\nodata \\\\ SBS1210+537A & 12h12m55.9s & +53d27m38s & 8989952 & 2004-06-06 & \\nodata & 480 & 240 \\\\ Tol1214-277 & 12h17m17.1s & -28d02m33s & 9008128 & 2004-06-28 & 0.0260 & 480 & 240 \\\\ Tol65 & 12h25m46.9s & -36d14m01s & 4829696 & 2004-01-07 & 0.0090 & 480 & 240 \\\\ UGCA292 & 12h38m40.0s & +32d46m01s & 4831232 & 2004-01-07 & 0.0010 & 480 & 240 \\\\ Tol1304-353 & 13h07m37.5s & -35d38m19s & 9006848 & 2004-06-25 & 0.0140 & 480 & 240 \\\\ Pox186 & 13h25m48.6s & -11d37m38s & 9007360 & 2004-07-14 & 0.0039 & 480 & 240 \\\\ CG0563 & 14h52m05.7s & +38d10m59s & 8992512 & 2005-05-30 & 0.0324 & 240 & 120 \\\\ CG0598 & 14h59m20.6s & +42d16m10s & 8992256 & 2005-03-19 & 0.0575 & 480 & 240 \\\\ CG0752 & 15h31m21.3s & +47d01m24s & 8991744 & 2005-03-19 & 0.0211 & 480 & 240 \\\\ Mrk1499 & 16h35m21.1s & +52d12m53s & 9011456 & 2004-06-05 & 0.0090 & 480 & 240 \\\\ {\\rm [RC2]}A2228-00 & 22h30m33.9s & -00d07m35s & 9006080 & 2004-06-24 & 0.0052 & 480 & 240 \\\\ \\enddata \\tablecomments{The coordinates and redshifts of the objects are cited from the NASA/IPAC Extragalactic Database (NED). In this paper, we only include the analysis of thirteen out of twenty-two sources which have SNRs sufficient for our abundance study. CG0563, CG0598 and CG0752 are included in the original sample as BCD candidates, however, they appear to be more starburst like \\citep[see][]{Hao07}. Thus even though they have high SNR, we do not include these three sources in this study.} \\end{deluxetable*} Metallicity is a key parameter that influences the formation and evolution of both stars and galaxies. Detailed studies of the elemental abundances of BCDs have already been carried out by several groups \\citep{Izotov99b, Kniazev03, Shi05} and the well known metallicity-luminosity relation has also been studied in detail in the environment of dwarf galaxies \\citep{Skillman89, Hunter99, Melbourne02}. However, because these studies were performed in the optical, they were limited by the fact that the properties of some of the deeply obscured regions in the star-forming galaxies may remain inaccessible due to dust extinction at these wavelengths. In fact, \\citet{Thuan99}, using ISO, have shown that the eastern component of SBS\\,0335-052 does have an embedded super star cluster (SSC) that is invisible in the optical while contributing $\\sim$75\\% to the bolometric luminosity \\citep[see also][]{Plante02,Houck04b}, even though it has very low metallicity (12+log(O/H)=7.33), which would in principle imply a low dust content. In addition to probing the dust enshrouded regions, emission in the infrared also has the advantage that the lines accessible at these wavelengths are less sensitive to the electron temperature fluctuations than the corresponding optical lines of the same ion. In the infrared, more ionization stages of an element become available as well. The improved sensitivity of the Infrared Spectrograph (IRS\\footnote{The IRS was a collaborative venture between Cornell University and Ball Aerospace Corporation funded by NASA through the Jet Propulsion Laboratory and the Ames Research Center.}) \\citep{Houck04a} on Spitzer has enabled us to obtain for the first time infrared spectra for a much larger sample of BCDs than was previously possible \\citep{Thuan99, Madden00, Verma03, Martin06}, thus motivating this study to probe the heavy element abundances in BCDs. In this paper, we analyze Spitze/IRS spectra of thirteen BCDs and present elemental abundances of neon and sulfur, which are both primary elements produced by the same massive stars in the nuclear synthesis processes. In section 2, we describe the sample selection, observations and data reduction. We present our results on the chemical abundances in section 3, followed by a comparison of the optical and infrared derived abundances in section 4. We show the interplay between the abundances and PAH emission in section 5. Finally, we summarize our conclusions in section 6. ", "conclusions": "We have studied the neon and sulfur abundances of thirteen BCDs using Spitzer/IRS high-resolution spectroscopy. Our analysis was based on the fine-structure lines and the hydrogen recombination line detected in the SH spectra of the {\\em IRS}, combined with the radio continuum, H$\\alpha$ images and integrated optical spectral data in some cases. We find a positive correlation between the neon and sulfur abundances, though sulfur appears to be more under-abundant than neon (with respect to solar). The ratio of Ne/S for our sources is on average 11.4$\\pm$2.9, which is consistent with what has been found in other HII regions using infrared data. However, this average ratio appears to be higher than the corresponding optical value of 6.5$\\pm$1.8 (in BCDs), which could be due to the adopted ICFs in the optical studies. When comparing the newly derived neon and sulfur abundances with the oxygen abundances measured from the optical lines, we find a good overall agreement. This indicates that there are few completely dust enshrouded HII regions in our BCDs, or if such HII regions are present, they have similar metallicities to the ones probed in the optical. Finally, the infrared derived neon and sulfur abundances also correlate, with some scatter, with the corresponding elemental abundances derived from the optical data." }, "0710/0710.0185.txt": { "abstract": "Observational and theoretical evidence suggests that coronal heating is impulsive and occurs on very small cross-field spatial scales. A single coronal loop could contain a hundred or more individual strands that are heated quasi-independently by nanoflares. It is therefore an enormous undertaking to model an entire active region or the global corona. %therefore requires simulating the %time-dependent plasma evolution of enormous numbers of %elemental strands. Three-dimensional MHD codes have inadequate spatial resolution, and 1D hydro codes are too slow to simulate the many thousands of elemental strands that must be treated in a reasonable representation. Fortunately, thermal conduction and flows tend to smooth out plasma gradients along the magnetic field, so ``0D models\" are an acceptable alternative. We have developed a highly efficient model called Enthalpy-Based Thermal Evolution of Loops (EBTEL) that accurately describes the evolution of the average temperature, pressure, and density along a coronal strand. It improves significantly upon earlier models of this type---in accuracy, flexibility, and capability. It treats both slowly varying and highly impulsive coronal heating; it provides the time-dependent differential emission measure distribution, $D\\!E\\!M(T)$, at the transition region footpoints; and there are options for heat flux saturation and nonthermal electron beam heating. EBTEL gives excellent agreement with far more sophisticated 1D hydro simulations despite using four orders of magnitude less computing time. It promises to be a powerful new tool for solar and stellar studies. ", "introduction": "An abundance of observational and theoretical evidence indicates that much of the corona is highly dynamic and evolves in response to heating that is strongly time-dependent. The evidence further suggests that the cross-field spatial scale of the heating is very small, so that unresolved structure is ubiquitous. In particular, many if not all coronal loops are bundles of thin strands that are heated impulsively and quasi-randomly by nanoflares. It is estimated that a single loop contains several tens to several hundreds of such strands. See \\citet{k06} for a detailed justification of these ideas and references to relevant work. Three-dimensional (3D) magnetohydrodynamic simulations are extremely useful for studying the source of coronal heating (instabilities of electric current sheets, reconnection, turbulence, etc.), but they cannot adequately address the complexity that is present in a single coronal loop, much less an entire active region. A more feasible approach is to treat the magnetic field as static and to solve the one-dimensional (1D) hydrodynamic equations along many representative flux strands using an assumed heating rate. The individual strands {\\it must} be treated separately. It is not valid to approximate a loop as a monolithic structure with uniform heating corresponding to the average for the component strands. This gives a completely different and incorrect result. There is reason to believe that the diffuse corona that lies between distinct bright loops is also comprised of elemental strands (e.g., \\citealt{anb07}). If roughly 100 strands are present in a single loop, then the numbers present in active regions and the global Sun are truly staggering. While it is possible to construct a limited number of model active regions with time-dependent 1D simulations \\citep{ww07}, it is not possible to investigate a wide range of values for the coronal heating parameters that must be assumed, such as the dependence on magnetic field strength, loop length, etc. \\citep{mdk00}. This is a major limitation, since we are still struggling to identify the properties and physical origin of the heating mechanism. Progress in the foreseeable future must therefore rely on simplified solutions to the hydro equations that treat field-aligned averages and are much less computationally intensive. These are sometimes called ``0D models\" because there is only one value of temperature, pressure, and density at any given time in the simulation. 0D models have been developed previously by \\citet{fh90} and \\citet{kp93}, but the best known is that of \\citet{c94}. It has been used to study a variety of topics, including coronal loops \\citep{ck97, ck06, kc01,petal06}, flares \\citep{rw02,pak02}, post-eruption arcades \\citep{rf05}, and active stellar coronae \\citep{ck06}. We have learned a great deal with the Cargill model, and our understanding has now advanced to the point where a more accurate and flexible model is required. This article presents an improved 0D model called Enthalpy-Based Thermal Evolution of Loops (EBTEL). As the name suggests, a key aspect of the model is an explicit recognition of the important role that enthalpy plays in the energy budget. EBTEL improves upon the Cargill model in several important ways. First, whereas the Cargill model is limited to an instantaneous heat pulse, EBTEL accommodates any time-dependent heating profile and can include a low-level background heating if desired. Second, EBTEL accounts for thermal conduction cooling and radiation cooling at all times during the evolution. The Cargill model assumes that only one or the other operates at any given time. Third, EBTEL has options for heat flux saturation and nonthermal electron beam heating. Finally, EBTEL is unique among 0D models in that it provides the time-dependent differential emission measure distribution of the transition region footpoints. Emission from the transition region plays a critical role in spatially unresolved observations, such as stellar observations and observations of the solar spectral irradiance, which is important for space weather \\citep{lean97}. Note that footpoint emission is not limited to the cooler ($<1$ MK) plasma traditionally associated with the transition region. It can also include hot emissions that originate from the base of very hot loops. The so-called moss seen in the ``coronal\" channels of the {\\it Transition Region and Coronal Explorer (TRACE)} is an example \\citep{betal99,mkb00}. We describe the coronal and transition region parts of EBTEL in the next two sections. We then present example simulations and compare them with corresponding simulations from a 1D model and, in one case, the Cargill model. We conclude with a discussion of EBTEL and the possible significance of the example simulations. ", "conclusions": "As evidenced by these examples, our simple 0D model is an excellent proxy for more sophisticated and far more computationally intensive 1D hydro simulations. It improves substantially on the 0D models of \\citet{c94}, \\citet{fh90}, and \\citet{kp93}. The Cargill model assumes that heating is instantaneous and that cooling occurs either by thermal conduction or by radiation, but not by both at the same time. The Fisher-Hawley model: (1) predicts abrupt evolutionary changes as the strand evolves between three distinct regimes; (2) does not account for the evaporation that continues well beyond the end of an impulsive heating event; and (3) cannot return to the pre-event state due to unphysical catastrophic cooling. The Kopp-Poletto model shares some similarities with EBTEL, but it treats the flows in a fundamentally different way. Like EBTEL, it equates the enthalpy carried by evaporative upflows with an excess heat flux, but the excess is determined relative to the pre-event state, rather than to the time-varying radiative losses from the transition region. Condensation downflows in the model are given by a density-dependent fraction of the free-fall velocity. In actuality, gravity plays no direct role in condensation, since the downflows are driven by pressure gradient deficits relative to hydrostatic equilibrium, in the same way that evaporative upflows are driven by pressure gradient excesses. Gravity sets the value of the hydrostatic gradient, but it is only the deficit or excess relative to this value that is important for the flows. Inclined strands experience essentially the same condensation and evaporation as do upright strands of the same length. Finally, EBTEL has advantages over all three of the other models in that it provides the $D\\!E\\!M(T)$ of the transition region and treats nonthermal electron beams and heat flux saturation. One obvious application of EBTEL is to investigate the idea that the basic structural elements of the corona are very thin, spatially unresolved magnetic strands that are heated impulsively. Loops may be bundles of such strands, as reviewed in \\citet{k06}, and the diffuse corona may be similarly structured. Differential emission measure distributions are one important test of this idea. Observed $D\\!E\\!M(T)$ curves from active regions and the quiet Sun tend to be peaked near $10^{6.5}$ and $10^{6.1}$ K, respectively, and to have a slope (temperature power law index) $\\ge 0.5$ coolward of the peak \\citep{rd81,dm93,betal96}. This is consistent with the coronal $D\\!E\\!M(T)$ curves of examples 1 and 4 (Figures \\ref{fig:dem1tr} and \\ref{fig:dem4}). The full loop curves are discrepant, on the other hand, due to the strong contribution from the transition region. The cited observations were made on the disk and should in principle include the transition region component. However, it is possible that absorption from chromospheric material such as spicules significantly attenuates the intensities of transition region lines used to construct the $D\\!E\\!M(T)$ curves (e.g., \\citealt{ddg95}; \\citealt{detal99}; \\citealt{df82}; \\citealt{so79}). We are currently investigating the magnitude of this effect. One of the great mysteries of coronal physics that has come to light in the last few years is the discovery that warm ($\\sim 1$ MK) coronal loops are much denser than expected for quasi-static equilibrium and live for much longer than a cooling time. The loops are therefore neither steadily heated nor cooling as monolithic structures. It has been shown that the observed densities and timescales can be explained by bundles of nanoflare heated strands, as long as nanoflares do not all occur at the same time (see \\citealt{k06} and references cited therein). Neighboring strands will therefore have different temperatures, and loops are predicted to have multi-thermal cross sections. In particular, emission should be produced at temperatures hotter than 3 MK. Hot loops are sometimes observed at the locations of warm loops, but not always. Example 5 suggests that nonthermal electron beams are a possible explanation for the lack of hot emission. As we have discussed, beams can produce excess densities through evaporation without the need for high temperatures. We have just begun to explore this possibility. For now, we note that the coronal $D\\!E\\!M(T)$ curve of example 5 (Figure \\ref{fig:dem5}) bears a close resemblance to the observed curves reported by \\citet{setal01} for a loop seen above the limb. %Other problems now being pursued with EBTEL include modeling the %emission characteristics of coronal arcades \\citep{pk07b}, modeling the %light curves of solar flares \\citep{rgm07}, investigating the %resolvability of elemental loop strands \\citep{pk07a}, and modeling coronal %loops as self-organized critical systems \\citep{lk05,kld06}. In this %last study, loop strands are assumed to become tangled by turbulent %photospheric %convection and to release magnetic energy when the misalignment between %adjacent strands reaches a critical angle, as expected for the secondary %instability \\citep{dka05}. EBTEL is used to follow the %plasma evolution of the strands. Preliminary results suggest that %this model can reproduce the light curves of loops observed by the %Soft X-ray Imager (SXI) on the GOES-12 satellite \\citep{lkm07}. However, %many more simulations with different combinations of model parameters are %necessary before any definitive conclusions can be drawn. This would %be very difficult with a computationally intensive 1D hydro code, but is %no problem with EBTEL. %Another major program we have begun under the partial sponsorship of %NASA\u0092s Living With a Star Program is to build realistic %models of active regions and the global Sun. Our approach is to %construct a ``magnetic skeleton\" %by extrapolating photospheric magnetograms and to populate many representative %field lines with plasma using EBTEL. Models of this kind have been %published before, but they use static equilibrium strand solutions and cannot %adequately reproduce imaging observations over %a wide range of temperatures \\citep{letal04, setal04, metal05, ww06, bw07}. %When the models are adjusted to %resemble soft X-ray images, they %they predict too little warm ($\\sim 1$ MK) emission at coronal altitudes %(EUV loops) and too much warm emission at the transition %region footpoints of hot coronal structures (EUV moss). There is %reason to believe that impulsive heating can improve the situation %\\citep{kld06, pk07b}. Indeed, \\citet{ww07} have built an impressive %model of an active region using the NRLFTM 1D hydro code and %find that impulsive heating gives a significant improvement over %steady heating. The agreement with observations is still not adequate, %however. %Since 1D simulations are computationally very intensive, %Warren and Winebarger were only able to consider one parameterized form for the %coronal heating function. Using EBTEL, we will examine a wide range of %heating parameters, including %the magnitude, duration, and frequency of the nanoflares, and their dependence %on magnetic field strength and field line length %(see \\citealt{mdk00}). We will also consider different proportions of %direct heating and nonthermal electron beams. %We estimate that 100 million hydro simulations %but is a manageable task with EBTEL. Identifying the coronal heating %parameters will provide valuable constraints for testing competing %theories of the heating mechanism. It should ultimately lead to a %physics-based operational model for nowcasting and %forecasting the X-ray and UV spectral irradiance. In conclusion, EBTEL is a powerful new tool that can be applied to a variety of problems in which large numbers of evolving strands must be computed. For example, it is now feasible to construct multiple models of nanoflare-heated active regions or entire stars and therefore to examine a wide array of nanoflare parameters (magnitude, lifetime, occurrence rate, dependence on field strength and strand length, etc.). By determining which parameters best reproduce the observations, we can place important constraints on the heating and thereby gain insight into the physical mechanism (e.g., \\citealt{mdk00}; \\citealt{setal04}; \\citealt{ww06}). EBTEL is currently being used to study the emission characteristics of coronal arcades \\citep{pk07b}, to explain the light curves of solar flares \\citep{rgm07}, and to model coronal loops as self-organized critical systems \\citep{lk05,kld06}. Interested users are invited to contact us for a copy of our IDL code. %We end by noting possible future improvements: %the inclusion of kinetic energy, cross-sectional area variation, and %gravitational stratification. %% Included in this acknowledgments section are examples of the %% AASTeX hypertext markup commands. Use \\url without the optional [HREF] %% argument when you want to print the url directly in the text. Otherwise, %% use either \\url or \\anchor, with the HREF as the first argument and the %% text to be printed in the second." }, "0710/0710.0373_arXiv.txt": { "abstract": "The physics behind the acceleration of the cosmic expansion can be elucidated through comparison of the predictions of dark energy equations of state to observational data. In seeking to optimize this, we investigate the advantages and disadvantages of using principal component analysis, uncorrelated bandpowers, and the equation of state within redshift bins. We demonstrate that no one technique is a panacea, with tension between clear physical interpretation from localization and from decorrelated errors, as well as model dependence and form dependence. Specific lessons include the critical role of proper treatment of the high redshift expansion history and the lack of a unique, well defined signal-to-noise or figure of merit. ", "introduction": "} The acceleration of the universe poses a fundamental mystery to cosmology, gravitation, and quantum physics. Understanding the nature of the dark energy responsible for the acceleration relies on careful, robust measurements of the dark energy properties, in particular its equation of state (EOS), or pressure to energy density, ratio that directly enters the Friedmann equation for cosmic acceleration. As scientists design the next generation of dark energy experiments they seek to optimize the measurements for the clearest insight into this unknown physics. Two critical pieces of information will be the value of the EOS at some epoch, such as the present, and a measure of its time variation, in much the way that early universe inflation theories are classified by the value of the spectral index and its running. The best parametrized EOS are physics based and model independent, i.e.\\ able to describe dark energy dynamics globally, or at least over a wide range of behaviors. Such EOS are very successful at fitting to data and projecting the results of future experiments, and can be robust to bias against inexact parametrization. Other approaches seek to remove one drawback of parametrized EOS by not assuming a functional form for the time variation, lest the true dark energy model lie outside the apparently wide range of validity of the form, i.e.\\ they aim for form independence. Two major avenues for achieving this are decomposition into basis functions or principal components (e.g.\\ \\cite{HutStark03}, also see \\cite{CritPog05,ShapTurn06,SimpBrid06,DickKnoxChu06,Ste06,HutPeir07}) and individual values of the EOS $w(z)$ over finite redshift bins, which become more general as the number of elements increases. However uncertainties in estimation of the EOS properties also grow as the number of principal components or bins increases. This article begins by examining general properties of the cosmological data and its dependence on the EOS in \\S\\ref{sec:cosdep}. Many of the later, detailed results will already be foreshadowed by this straightforward and general analysis. In \\S\\ref{sec:pca} we examine principal component analysis of the EOS and in \\S\\ref{sec:band} uncorrelated bandpowers. Bins of EOS in redshift is investigated in \\S\\ref{sec:bin}, including figures of merit for quantifying the uncertainties. Further concentration on the crucial role of the high redshift EOS, and the risk of biasing parameter estimation, occurs in \\S\\ref{sec:whi}. We consider physical constraints on EOS properties in \\S\\ref{sec:wlimit} and summarize our results and conclude in \\S\\ref{sec:concl}. ", "conclusions": "} The dark energy equation of state properties contain clues crucial to understanding the nature of the acceleration of the cosmic expansion. Deciphering those properties from observational data involves a combination of robust analysis and clear interpretation. We considered three approaches -- principal components, uncorrelated bandpowers, and binning; none of the approaches provides a panacea. In particular, we identify issues of dependence on basis functions, binning variables, and baseline models. The three approaches are not truly nonparametric and physical interpretation (not merely the values) of the results in the two decorrelated basis techniques depends on model, priors, and data, indeed even on an implicitly assumed functional form. Nevertheless, principal components can give a useful guide to the qualitative sensitivity, the best constrained aspects, of the data. The uncorrelated bin approach unfortunately does not truly deliver uncorrelated bandpowers for the equation of state. This approach using the square root of the Fisher matrix does not tightly localize the information (without a strong prior), making the interpretation nontrivial. This property of nonlocality is inherent in the cosmological characteristics. One might prefer to stay with the original binned equations of state used as the initial step for this technique, which are readily interpreted. Conversely, if the modes can be localized, the interpretation is easy, but in that case the original Fisher matrix is close to diagonal and thus the original bins almost uncorrelated. Hence, again, one might as well stay with the bin parameters which have a clear meaning. Indeed the goal is understanding the physics, not obtaining particular statistical properties. Decorrelated parameters that are not readily interpretable physically are of limited use; for example one still prefers to analyze the cosmic microwave background in terms of physical quantities such as physical matter density and spectral tilt rather than the principal axes of the eigenvectors. Note that the uncertainty on the EOS behavior $\\sigma(w(z))$ is the same whether calculated by PCA (if all modes are kept), uncorrelated bands, or binned EOS, since the same information is in the data. We also emphasize that the modes most clearly determine the effect on the equation of state, not the weights, which are often the only quantity displayed. Moderately localized, even all positive, weights do not guarantee a localized physical effect. A further caution is that locality and positivity of weights can owe more to prior restrictions, especially the treatment of the high redshift equation of state, than to the data itself. Assuming a fixed value for the high redshift equation of state has major, widespread impacts on the results, ranging from strongly misestimated uncertainties to spurious localization to bias in the derived cosmology. We emphasize that it is essential to fit for the high redshift behavior in order not to be misled. Adding CMB data and marginalizing over a new, high redshift bin removes the ill effects of bias but ``cancels out'', providing no new constraints; multiple data points for $z>2$ are required, such as from high redshift distances or weak lensing measurements of the mass growth behavior. Assuming that dark energy is negligible at $z>2$ is also effectively assuming a functional form -- precisely what the use of eigenmodes was supposed to avoid. Indeed, functional forms do not have many of the basis, model, binning, etc.\\ dependences of eigenmodes, while principal components are in turn not fully form independent. If one assumes a functional form to obtain informative constraints on the equation of state, one must indeed choose the form to represent robustly the physical behavior (as has been shown to be widely the case for $w(a)=w_0+w_a(1-a)$ by \\cite{Linder03,Linder0708}), and carefully check the range of validity of the conclusions by examining other forms. A good complementary analysis tool would be the binned equation of state approach examined here. Regardless of the form of analysis, only a finite amount of information can be extracted from even next generation data. As has been concluded for functional equations of state and principal component analysis \\cite{LinHut05}, the analysis here in terms of binned equation of state indicates that only two physically informative parameters can be fit with realistic accuracy. However, we identify several issues in the PCA and uncorrelated bin approaches that cause accuracy or signal to noise criteria to be ill defined. Similar difficulties arise in condensing the physical information on dark energy to a single figure of merit; the number is quite sensitive to cosmologically irrelevant aspects like the binning used (as well as very dependent on the treatment of the high redshift dark energy behavior). In conclusion, physically motivated fitting of the equation of state such as the $w_0$-$w_a$ parametrization in complement with a binned equation of state approach (perhaps with physical constraints such as outlined in \\S\\ref{sec:wlimit}) have the best defined, clearest to interpret, and robust insights of the approaches we considered. With any method, one must use caution regarding the influence of priors and fit the dark energy physics over the entire expansion history." }, "0710/0710.0780_arXiv.txt": { "abstract": "$Chandra$ and $XMM-Newton$ observations of the Cartwheel galaxy show $\\sim{}17$ bright X-ray sources ($\\gtrsim{}5\\times{}10^{38}$ erg s$^{-1}$), all within the gas-rich outer ring. We explore the hypothesis that these X-ray sources are powered by intermediate-mass black holes (IMBHs) accreting gas or undergoing mass transfer from a stellar companion. To this purpose, we run $N$-body/SPH simulations of the galaxy interaction which might have led to the formation of Cartwheel, tracking the dynamical evolution of two different IMBH populations: halo and disc IMBHs. Halo IMBHs cannot account for the observed X-ray sources, as only a few of them cross the outer ring. Instead, more than half of the disc IMBHs are pulled in the outer ring as a consequence of the galaxy collision. However, also in the case of disc IMBHs, accretion from surrounding gas clouds cannot account for the high luminosities of the observed sources. Finally, more than 500 disc IMBHs are required to produce $\\lesssim{}15$ X-ray sources via mass transfer from very young stellar companions. Such number of IMBHs is very large and implies extreme assumptions. Thus, the hypothesis that all the observed X-ray sources in Cartwheel are associated with IMBHs is hardly consistent with our simulations, even if it is still possible that IMBHs account for the few ($\\lesssim{}1-5$) brightest ultraluminous X-ray sources (ULXs). ", "introduction": "Ring galaxies are among the most fascinating objects in the Universe. The Cartwheel galaxy is surely the most famous of them, and also the biggest (with its optical diameter of $\\sim{}$50-60 kpc) and the most studied. Cartwheel has been thoroughly observed in almost every band: H$\\alpha{}$ (Higdon 1995) and optical (Theys \\& Spiegel 1976, Fosbury \\& Hawarden 1977) images, red continuum (Higdon 1995), radio line (Higdon 1996) and continuum (Higdon 1996; Mayya et al. 2005), near- (Marcum, Appleton \\& Higdon 1992) and far-infrared (Appleton \\& Struck-Marcell 1987a), line spectroscopy (Fosbury \\& Hawarden 1977) and X-ray (Wolter, Trinchieri \\&{} Iovino 1999; Gao et al. 2003; Wolter \\& Trinchieri 2004; Wolter, Trinchieri \\& Colpi 2006). Cartwheel exhibits a double ringed shape, with some transversal 'spokes' (Theys \\& Spiegel 1976, Fosbury \\& Hawarden 1977; Higdon 1995), which have been detected only in few of the $\\lesssim{}300$ known ring galaxies (Arp \\& Madore 1987; Higdon 1996). The outer ring is rich of HII regions, especially in its southern quadrant, and dominates the H$\\alpha{}$ emission (Higdon 1995). This implies that Cartwheel is currently undergoing an intense epoch of star formation (SF, with SF rate $\\sim{}20-30\\,{}M_\\odot{}$ yr$^{-1}$; Marston \\& Appleton 1995; Mayya et al. 2005), almost entirely confined to the outer ring. Inner ring, nucleus and spokes are nearly devoid of gas and dominated by red continuum emission, indicating a relatively old, low- and intermediate-mass stellar population (Higdon 1995, 1996; Mayya et al. 2005). Cartwheel is located in a small group of 4 galaxies. All of its 3 companions (known as G1, G2 and G3) are less massive than Cartwheel, and host less than 20 per cent of the total gas mass in the group. The analysis of X-ray data shows another intriguing peculiarity: most of the point sources detected by $Chandra$ are in the outer ring, and particularly concentrated in the southern quadrant (Gao et al. 2003; Wolter \\& Trinchieri 2004). According to Wolter \\& Trinchieri (2004) 13 out of 17 sources associated with Cartwheel are in the outer ring, the remaining 4 being close to the inner rim of the ring or to the optical spokes\\footnote{The total number of point X-ray sources in the Cartwheel group is 24 (Wolter \\& Trinchieri 2004), but 6 of them are associated with the companion galaxies G1 and G2, whereas 1 of them is probably background/foreground contamination.}. Gao et al. (2003) noted that all the five strongest H$\\alpha{}$ knots in the ring are associated with an X-ray source, indicating a possible correlation between X-ray sources and young star-forming regions. Furthermore, most of the observed sources have isotropic X-ray luminosity $L_X\\gtrsim{}10^{39}$ erg s$^{-1}$, belonging to the category of ultraluminous X-ray sources (ULXs). Given the distance of Cartwheel ($\\sim{}124$ Mpc for an Hubble constant $H_0=73$ km s$^{-1}$ Mpc$^{-1}$), the data are incomplete for $L_X\\lesssim{}5\\times{}10^{38}$ erg s$^{-1}$, and the faintest sources in the sample have $L_X\\sim{}10^{38}$ erg s$^{-1}$. Then, almost all the X-ray sources detected in Cartwheel are close to the ULX range. Many theoretical studies, both analytical (Lynds \\& Toomre 1976; Theys \\& Spiegel 1976; Appleton \\& Struck-Marcell 1996) and numerical (Theys \\& Spiegel 1976; Appleton \\& Struck-Marcell 1987a, 1987b; Hernquist \\& Weil 1993; Mihos \\& Hernquist 1994; Struck 1997; Horellou \\& Combes 2001; Griv 2005) were aimed to explain the origin of Cartwheel and its observational features. In the light of these papers, the origin of propagating rings in Cartwheel and similar galaxies can be explained by galaxy collisions with small impact parameter (Theys \\& Spiegel 1976; Appleton \\& Struck-Marcell 1987a, 1987b; Hernquist \\& Weil 1993; Mihos \\& Hernquist 1994; Struck 1997; Horellou \\& Combes 2001). Among the companions of Cartwheel, both G1 and G3 are good candidates for this interaction, having a small projected impact parameter, being not too far from Cartwheel, and showing a disturbed distribution of neutral hydrogen (Higdon 1996). An alternative, less investigated model explains the rings with disc instabilities (Griv 2005). Models of galaxy collisions explain quite well most of Cartwheel properties. However, none of the previous theoretical studies has investigated the nature of the X-ray sources, and especially of the ULXs, observed in Cartwheel. The nature of the ULXs is still not understood. It has been suggested that they are high-mass X-ray binaries (HMXBs) powered by stellar mass black holes (BHs) with anisotropic X-ray emission (King et al. 2001; Grimm, Gilfanov \\& Sunyaev 2003; King 2006) or with super-Eddington accretion rate (e.g. King \\&{} Pounds 2003; Socrates \\& Davis 2006; Poutanen et al. 2007). However, some ULXs, especially the brightest ones ($L_X\\gtrsim{}10^{40}$ erg s$^{-1}$), show characteristics which are difficult to reconcile with the hypothesis of beamed emission, such as the presence of an isotropically ionized nebula (e.g. Pakull \\& Mirioni 2003; Kaaret, Ward \\& Zezas 2004) or of quasi periodic oscillations (e.g. Strohmayer \\& Mushotzky 2003). Then, it has been proposed that some ULXs (or at least the brightest among them; Miller, Fabian \\& Miller 2004) might be associated with intermediate-mass black holes (IMBHs), i.e. BHs with mass $20\\lesssim{}m_{\\rm BH}/M_\\odot{}\\lesssim{}10^{5}$, accreting either gas or companion stars (Miller et al. 2004; see Mushotzky 2004 and Colbert \\& Miller 2005 for a review). On the other hand, most of ULXs can be explained with the properties of stellar mass BHs (see Roberts 2007 for a review and references therein). In this paper we present a new, refined $N-$body/SPH model of the Cartwheel galaxy, which reproduces the main features of Cartwheel. The aim of this paper is to use the $N-$body/SPH model in order to check whether the IMBH hypothesis is viable to explain all or a part of the X-ray sources observed in Cartwheel. In fact, in the last few years the hypothesis that most of the ULXs are powered by IMBHs has became increasingly difficult to support, as the observational features of the majority of ULXs are consistent with accreting stellar mass BHs (Roberts 2007). It would be interesting to see whether also the dynamical simulations agree with this conclusion\\footnote{We will not consider other possible scenarios (e.g. the production of ULXs by beamed emission in HMXBs), due to intrinsic limits of $N-$body methods.}. In Section 2 we describe our simulations. In Section 3 we discuss the evolution of our models and the dynamics of either halo or disc IMBHs. In Section 4 we investigate the possibility that the X-ray sources in Cartwheel are powered by IMBHs accreting gas or stars. In the last case we consider both the accretion from old stars (which produce only transient sources) and from young stars, formed after the galaxy collision (which can lead also to persistent sources). We also investigate the hypothesis that new disc IMBHs are formed in very young clusters after the galaxy collision. Our findings are summarized in Section 5. ", "conclusions": "In this paper we investigated the possible connection among IMBHs and the $\\sim{}17$ (Gao et al. 2003; Wolter \\& Trinchieri 2004) bright X-ray sources detected in the outer ring of Cartwheel. Recent observations show that models based on beamed emission or super-Eddington accretion in HMXBs including stellar mass BHs can explain most of ULXs, apart from the brightest ones (Roberts 2007 and references therein). However, the observations cannot definitely exclude that all the ULXs are powered by IMBHs. Thus, in our paper we checked whether the IMBHs can account for all or only for a part of the ULXs observed in Cartwheel. We simulated the formation of a Cartwheel-like ring galaxy via dynamical interaction with an intruder galaxy. In this simulation we also integrated the evolution of 100 IMBHs particles. We considered two different models of IMBH formation, i.e. IMBHs born as relics of population III stars (and distributed as a concentrated halo population) and IMBHs formed via runaway collapse of stars (and distributed as an exponential disc). For these models, we investigated both gas accretion from surrounding molecular clouds and mass transfer from a stellar companion. The main results of this study are the following. \\subsection{Halo IMBHs} IMBHs born as the relics of population III stars, if they are distributed as a halo population, cannot contribute to the X-ray sources, neither via gas accretion nor via mass transfer in binaries. In particular, the luminosity produced by halo IMBHs accreting gas is always many orders of magnitude smaller than that of the observed sources, even if we assume that halo IMBHs have a large mass ($m_{\\rm BH}=10^3\\,{}M_\\odot{}$) and a high radiative efficiency ($\\eta{}=0.1$). This is due to the fact that only a small fraction of halo IMBHs passes through the disc (only for a short lapse of time), and even these IMBHs have a high ($v\\gtrsim{}100$ km s$^{-1}$) relative velocity with respect to gas particles. Similarly, halo IMBHs cannot accrete mass from stars, as the probability that they acquire a companion is very low. \\subsection{Disc IMBHs} IMBHs born from the runaway collapse of stars should be a disc population. Under overoptimistic assumptions ($n_g=10^2\\,{}{\\rm cm}^{-3}$, $m_{\\rm BH}=10^3\\,{}M_\\odot{}$ and $\\eta{}=0.1$) these IMBHs can produce, via gas accretion, X-ray sources with $L_X\\lesssim{}10^{39}$ erg s$^{-1}$, a factor of $\\sim{}10$ fainter than the brightest ULXs observed in Cartwheel. Thus, also disc IMBHs accreting gas cannot explain the observed X-ray sources in Cartwheel. Our model of gas accreting IMBHs contains many rough assumptions; but most of them are upper limits, strengthening our result. However, a more realistic treatment of the local properties of the gas would be helpful, to understand the physical mechanisms of gas accretion onto IMBHs. On the other hand, runaway collapse born IMBHs are hosted in dense young clusters. In such environment, it is easy for the IMBH to capture a stellar companion. Blecha et al. (2006) estimated that a $\\sim{}100\\,{}M_\\odot{}$ IMBH undergoes mass transfer from a companion star for about the $3$ per cent of the cluster lifetime. Previous papers (Portegies Zwart et al. 2004; Patruno et al. 2005) have shown that IMBHs accreting from low-mass ($\\lesssim{}10\\,{}M_\\odot{}$) and high-mass ($\\gtrsim{}10\\,{}M_\\odot{}$) companions generate transient (with a bright phase of only a few days) and persistent bright X-ray sources, respectively. As $10\\,{}M_\\odot{}$ stars have a lifetime of $\\sim{}30-40\\,{}$ Myr, only IMBHs hosted in sufficiently young star clusters can generate persistent X-ray sources. Then, IMBHs hosted in clusters born before the dynamical encounter with the intruder (i.e. more than 100 Myr ago) can produce only transient sources. We estimated that, out of 100 IMBHs which were present before the dynamical encounter with the intruder galaxy, only $\\lesssim{}2-3$ are expected to undergo mass transfer from low-mass companions at present, producing a comparable number of transient X-ray sources. As observations show that at least 4 X-ray sources in the Cartwheel ring are persistent over a time-scale of 6 months (Wolter et al. 2006), we conclude that pre-encounter formed IMBH binaries are not sufficient to explain the data. We considered the possibility that pre-encounter IMBHs capture massive stars produced after the encounter with the intruder. In this case, under overoptimistic assumptions, 100 pre-encounter IMBHs can produce $\\lesssim{}1$ X-ray source, either persistent or not. Finally, we hypothesized that very young ($<40$ Myr) star clusters, formed after the encounter, generate IMBHs at their centre. Under this hypothesis, $500-1000$ IMBHs are required to produce $\\sim{}15-30$ bright ($10^{36}-10^{41}$ erg s$^{-1}$) X-ray sources, some of them persistent and some transient% . This scenario might account for the $\\sim{}17$ observed X-ray sources in the Cartwheel ring. It is also in agreement with the fact that many ULXs observed in Cartwheel are associated with bright $H_\\alpha{}$ spots, i.e. active star-forming regions (Gao et al. 2003). The birth of $\\sim{}1000$ IMBHs (each one of 100 $M_\\odot{}$) in $\\sim{}$40 Myr implies an IMBH formation rate of $2.5\\times{}10^{-3}\\,{}M_\\odot{}\\,{}{\\rm yr}^{-1}$, that is a factor of $\\sim{}10^4$ lower than the SF rate. This rate is acceptable for runaway collapse scenarios, as Portegies Zwart \\& McMillan (2002) show that $\\sim{}$0.1 per cent of the mass of the parent young cluster merges to form the IMBH. However, we stress that only the runaway collapse scenario, under extreme assumptions, can explain the formation of such a huge number of IMBHs in the disc. Thus, our simulations suggest that IMBHs can hardly account for all the ULXs observed in Cartwheel. On the other hand, it is possible that only the few brightest sources in Cartwheel are powered by IMBHs, while the other ones are either beamed HMXBs, super-Eddington accreting stellar mass BHs or a blending of multiple fainter sources. For example, $\\sim{}30$ IMBHs are expected to form in the ring, in order to produce only 1 very bright ULX, such as the source N.10 in Cartwheel. These results agree with the semi-analytical model by King (2004), who showed that IMBHs cannot explain all the ULXs in Cartwheel. However, King (2004) concludes that $>3\\times{}10^4$ IMBHs are required to produce the observed number of ULXs, $\\sim{}30-60$ times more than in our analysis. This apparent discrepancy is due to the fact that King (2004) assumes that the IMBHs power only transient sources (Kalogera et al. 2004), and thus he has to introduce a $\\sim{}10^{-2}$ duty-cycle. However, Patruno et al. (2005) showed that IMBHs accreting from young massive stars ($\\gtrsim10\\,{}M_\\odot{}$) produce non-transient sources, increasing the expected duty-cycle. In conclusion, new $Chandra$ and $XMM-Newton$ observations of Cartwheel could partially solve the mystery of Cartwheel X-ray sources, investigating which sources are transient, which variable, and which persistent. Deeper observations are also needed to resolve possible blended sources. In the future, it would be interesting to search whether other ring galaxies host as many bright X-ray sources as Cartwheel and whether these sources are similarly concentrated in the outer ring." }, "0710/0710.0749_arXiv.txt": { "abstract": "{}{We investigate the structure of a field around the position of V838~Mon as seen in the lowest CO rotational transitions. We also measure and analyse emission in the same lines at the position of V838~Mon.}{Observations have primarily been done in the $^{12}$CO $J = 2$$\\to$$1$ and $J=3$$\\to$$2$ lines using the KOSMA telescope. A field of 3.4 squared degrees has been mapped in the on-the-fly mode in these transitions. Longer integration spectra in the on-off mode have been obtained to study the emission at the position of V838~Mon. Selected positions in the field have also been observed in the $^{12}$CO $J = 1$$\\to$$0$ transition using the Delingha telescope.}{In the observed field we have identified many molecular clouds. They can be divided into two groups from the point of view of their observed radial velocities. One, having $V_{\\rm LSR}$ in the range $18-32\\,{\\rm km\\,s}^{-1}$, can be identified with the Perseus Galactic arm. The other one, having $V_{\\rm LSR}$ between $44-57\\,{\\rm km\\,s}^{-1}$, probably belongs to the Norma-Cygnus arm. The radial velocity of V838~Mon is within the second range but the object does not seem to be related to any of the observed clouds. We did not find any molecular buble of a $1\\degr$ dimension around the position of V838~Mon claimed in van~Loon et~al. An emission has been detected at the position of the object in the $^{12}{\\rm CO}\\ J=2$$\\to$$1$ and $J=3$$\\to$$2$ transitions. The emission is very narrow (FWHM~$\\simeq$~1.2~km\\,s$^{-1}$) and at $V_{\\rm LSR} = 53.3$~km\\,s$^{-1}$. Our analysis of the data suggests that the emission is probably extended.}{} ", "introduction": "} The outburst of V838 Mon was discovered in the beginning of January~2002. Initially thought to be a nova, the object appeared unusual and enigmatic in its nature. The eruption, as observed in the optical, lasted about three months (Munari et~al. \\cite{muna02}, Kimeswenger et~al. \\cite{kimes02}, Crause et~al. \\cite{crause03}). After developing an A--F supergiant spectrum at the maximum at the beginning of February~2002, the object evolved to lower effective temperatures and in April~2002 it practically disappeared from the optical, remaining very bright in the infrared. A detailed analysis of the evolution of the object in the outburst and decline can be found in Tylenda (\\cite{tyl05}). Different outburst mechanisms have been proposed to explain the eruption of V838~Mon. They include an unusual nova (Iben \\& Tutukov \\cite{it92}), a late He-shell flash (Lawlor \\cite{law05}) and a stellar merger (Soker \\& Tylenda \\cite{soktyl03}). These models have critically been discussed in Tylenda \\& Soker (\\cite{tylsok06}) and the authors conclude that the only mechanism that can satisfactorily account for the observational data is a collision and merger of a low-mass pre-main-sequence star with an $\\sim$$8\\,M_\\odot$ main-sequence star. V838 Mon is surrounded by diffuse matter which gave rise to a spectacular light-echo phenomenon (e.g. Bond et~al. \\cite{bond03}). Bond et~al. claim that the matter comes from previous eruptions of the object. However Tylenda (\\cite{tyl04}), Crause et~al. (\\cite{crause05}), as well as Tylenda, Soker \\& Szczerba (\\cite{tss05}) argue that the echoing matter is of interstellar origin. The latter is consistent with recent findings, namely that V838~Mon is a member of a young open cluster (Afsar \\& Bond \\cite{afsar}) and that the total mass of the echoing matter is probably of the order of $100\\ M_\\odot$ (Banerjee et~al. \\cite{baner06}). van Loon et al. (\\cite{loon04}) have analyzed archive infrared and radio data on the sky around V838~Mon and claimed discovery of multiple shells ejected by the object in the past. In particular, from a compilation of CO galactic surveys done in Dame et~al. (\\cite{dame01}) van~Loon et~al. have suggested that V838~Mon is situated within a bubble of CO emission with a diameter of $\\sim$$ 1\\degr$. These results have been critically discussed in Tylenda et~al. (\\cite{tss05}), who have concluded that the shells of van~Loon et~al. are not realistic. Deguchi et~al. (\\cite{degu05}) have discovered an SiO maser emission from V838~Mon. The main component is narrow and centered at $V_{\\rm LSR} \\simeq 54\\ {\\rm km}\\,{\\rm s}^{-1}$, which is thought to be a radial velocity of the object itself. Further observations of Claussen et~al. (\\cite{clauss06}) have shown that the SiO maser is variable and that most of the emission comes from a region smaller than a milliarcsecond. In the present paper we report on results of our observations of V838~Mon and its nearby vicinity in the $^{12}$CO $J = 1$$\\to$0, 2 $\\to$1 and 3$\\to$2 transitions. We describe the observational material and discuss results on the CO emission from a field of 3.4 squared degrees around the position of V838~Mon. One of the goal of this survey is to verify the existence of the CO bubble around V838~Mon claimed in van~Loon et~al. (\\cite{loon04}). Measurements of the CO emission obtained at the position of V838~Mon are also presented, analysed and discussed. A preliminary analysis of the data has been done in Kami\\'{n}ski et~al. (\\cite{kmst}). ", "conclusions": "} Our on-the-fly maps in the CO rotational transitions, especially those done in the (2$\\to$1) line with the KOSMA telescope, have allowed us to identify 29 molecular clouds around the position of V838~Mon (see Appendix~\\ref{append}). As far as we know, the region mapped with the KOSMA telescope has not been observed before with a sensitivity and a spatial resolution comparable to our survey. The data presented in Appendix~\\ref{append} can be used in studies of molecular matter in the outer Galaxy. We do not confirm the existence of a molecular bubble of a 1 degree dimension claimed in van~Loon et~al. (\\cite{loon04}), which was probably an artefact resulted from merging two surveys of different resolution and sensitivity in the data used by van~Loon et~al. Our maps did not reveal any molecular cloud in the near vicinity of V838~Mon. The nearest CO emission has been detected at $\\sim$$40$~arcmin from the position of the object (see Fig.~\\ref{map_2}). Thus the nearest dense molecular cloud is located at least at a distance of $\\sim$$80$~pc from V838~Mon (assuming a 7~kpc distance to V838~Mon -- see below). From the noise level in the (2$\\to$1) map (see Table~\\ref{tech_tab}) we can put an upper limit to $I_{\\rm CO}=\\int T_{\\rm mb}\\,dV$ of $1.26~{\\rm K~km~s}^{-1}$ ($3 \\sigma_{\\rm rms}$) at the position of V838~Mon. Assuming the $N$(H$_2$) to $I_{\\rm CO}$ conversion factor of $X_{\\rm CO} = 5.1$ (in units $10^{20}\\,{\\rm molecules}\\ {\\rm cm}^{-2}\\,{\\rm K}^{-1}\\,{\\rm km}^{-1}\\,{\\rm s}$, typical value for the NC clouds discussed in Appendix~\\ref{append}) we obtain an upper limit to the column density of $N({\\rm H}_2) = 6.4 \\times 10^{20}~{\\rm cm}^{-2}$. This upper limit is an order of magnitude lower than typical column densities observed for molecular clouds in the vicinity of the Sun (e.g. Blitz \\cite{blitz}). Thus we can conclude that V838~Mon is not located in a typical molecular cloud. However, as presented in Sect.~\\ref{kosma}, long integrations in the on-off mode have allowed us to detect an emission in the CO (2$\\to$1) and (3$\\to$2) transitions at the position of V838~Mon. The question that arises is: what is the origin of this emission? Certainly it cannot come from the matter ejected in the 2002 outburst. That matter was expanding with large velocities, a few hundred km~s$^{-1}$, while our profiles have a FWHM of $\\sim$$1\\,{\\rm km\\,s}^{-1}$. The $V_{\\rm LSR}$ position and width of the CO lines similar to those of the SiO maser (Deguchi et~al. \\cite{degu05}, Claussen et~al. \\cite{clauss06}) suggest that both emissions (CO and SiO) originate from the same place, e.g from the remnant of the 2002 outburst. However, in this case the KOSMA telescope would see a point source, which does not seem to be the case. The April 2006 observations show detectable emissions at the 1~arcmin offsets from the V838~Mon position (see Table~\\ref{res2_tab}). The HPBW of the KOSMA beam at 230~GHz is 130~arcsec. If the CO emission source were a point source at the (0,0) position, than the intensity measured at the (0,1) and ($0,-1$) offsets would be about twice fainter than that at the central position. Table~\\ref{res2_tab} shows that this ratio is larger, i.e. $0.6-0.8$. Unfortunantely the accuracy of the measurements at the 1~arcmin offsets was not high, so we cannot say that the results in Table~\\ref{res2_tab} are definitively inconsistent with a point source emission. However, there are other arguments in favour of the idea that the CO emission is extended and/or situated outside the V838~Mon position. Our Delingha observations in the (1$\\to$0) transition (see Sect.~\\ref{delingha}) have not detected any emission at the V838~Mon position and the upper limit was 0.39~K. Assuming that the intensity in the (1$\\to$0) line is comparable to that in the (2$\\to$1) line (which is usually the case in molecular clouds) and that we observe a point source at the V838~Mon position than from the measured $T_{\\rm mb}$ in the (2$\\to$1) transition in December~2005 (see Table~\\ref{res_tab}), taking into account different beamsizes of the two telescopes (55~arcsec in Delingha versus 130~arcsec in KOSMA), the expected value of $T_{\\rm mb}$ in the (1$\\to$0) Delingha observations would be $\\sim$$1.8$\\,K. This is 4.5 times higher than the observed upper limit. Thus either the (2$\\to$1)/(1$\\to$0) ratio is exceptionally large ($> 4.5$) or the CO source is situated outside the Delingha beam but still inside the KOSMA beam. The later interpretation is supported by our possible detection of an emission with Delingha at three positions $\\sim$$1$\\,arcmin from the object (see Table~\\ref{resd_tab}). It is also supported by a finding of Deguchi et~al. (\\cite{degu06}) who, using the Nobeyama telescope, tried to measure the CO (1$\\to$0) emission from a few points in a field around V838~Mon. A signal was detected from a position 30~arcsec north of V838~Mon at a velocity very close to that of the (2$\\to$1) line in our KOSMA observations. No emission was however recorded at the position of the object. The above discussion allows us to conclude that the CO emission, clearly seen in our KOSMA observations, most probably originates not from V838~Mon itself but from a region (regions) situated, typically, $\\sim$$1$\\,arcmin from the V838~Mon position. There are two possible ways of explaining the emission in this case. One is that the observed emission is a faint part of a larger CO structure belonging, for instance, to a molecular cloud. However, as discussed above, our CO maps have not revealed any dense CO cloud of similar $V_{\\rm LSR}$ closer than $\\sim$$40$\\,arcmin from the position of V838~Mon. All the detected and possibly detected CO emission at and near the position of V838~Mon lie well within the optical light echo of V838~Mon. Hence the second possibility, namely that the CO emission comes from the same matter that gave rise to the light echo and that it was the light flash from the 2002 eruption which induced the emission. Then the observed narrowness of the line profile would be a strong argument that the matter is of interstellar origin rather than being ejected by V838~Mon in previous eruptions. The observed $V_{\\rm LSR} = 53.3\\,{\\rm km\\,s}^{-1}$ of the CO emission would imply a kinematical distance of $\\sim$$7.0$\\,kpc (using the Galactic rotational curve of Brandt \\& Blitz \\cite{bb93}). This can be compared to $6.1 - 6.2$~kpc found by Bond et~al. (\\cite{bond06}) and $\\sim$$8$~kpc advocated in Tylenda (\\cite{tyl05}). Rushton et al. (\\cite{rush03}) searched for CO emission from V838~Mon about a year after the 2002 eruption. No measurable signal was detected in the three lowest rotational transitions and the upper limit was $T_A^* \\la 25-40$~mK. It is not straightforward to compare this result with our findings as the observations of Rushton et~al. were done about 3~years before ours and the object probably evolved significantly during this time span. Nevertheless, given the beamwidth of the telescopes used by Rushton et~al. (HPBW~$\\le 45\\,\\arcsec$), their negative result at the position of the object does not preclude a possibility that a significant emission was present in a near vicinity of the object, but outside the telescope beam. More sensitive observations with a higher spatial resolution are required to distinguish between the above discussed possibilities and to futher investigate the problem of the CO emission from V838~Mon." }, "0710/0710.5393_arXiv.txt": { "abstract": "The Zeeman pattern of Mn~{\\sc i} lines is sensitive to hyperfine structure (HFS) and, because of this reason, they respond to hG magnetic field strengths differently from the lines commonly used in solar magnetometry. This peculiarity has been employed to measure magnetic field strengths in quiet Sun regions. However, the methods applied so far assume the magnetic field to be constant in the resolution element. The assumption is clearly insufficient to describe the complex quiet Sun magnetic fields, biasing the results of the measurements. The diagnostic potential of Mn~{\\sc i} lines can be fully exploited only after understanding the sense and the magnitude of such bias. We present the first syntheses of Mn~{\\sc i} lines in realistic quiet Sun model atmospheres. Plasmas varying in magnetic field strength, magnetic field direction, and velocity, contribute to the synthetic polarization signals. The syntheses show how the Mn~{\\sc i} lines weaken with increasing field strength. In particular, kG magnetic concentrations produce \\mni{5538} circular polarization signals (Stokes~$V$) which can be up to two orders of magnitude smaller than what the weak magnetic field approximation predicts. As corollaries of this result, (1) the polarization emerging from an atmosphere having weak and strong fields is biased towards the weak fields, and (2) HFS features characteristic of weak fields show up even when the magnetic flux and energy are dominated by kG fields. For the HFS feature of \\mni{5538} to disappear the filling factor of kG fields has to be larger than the filling factor of sub-kG fields. Since the Mn~{\\sc i} lines are usually weak, Stokes~$V$ depends on magnetic field inclination according to the simple consine law, a scaling that holds independently of the magnetic field strength. Atmospheres with unresolved velocities produce very asymmetric line profiles, which cannot be reproduced by simple one-component model atmospheres. Inversion techniques incorporating complex magnetic atmospheres must be implemented for a proper diagnosis. Using the HFS constants available in the literature we reproduce the observed line profiles of nine lines with varied HFS patterns. Consequently, the uncertainty of the HFS constants do not seem to limit the use of Mn~{\\sc i} lines for magnetometry. ", "introduction": "\\label{intro} When the polarimetric sensitivity and the angular resolution exceed the required threshold, magnetic signals show up almost everywhere on the solar photosphere. The signals in supergranulation cell interiors are particularly weak, however, most of the solar surface is in the form of cell interiors and, therefore, these weak signals probably set the total unsigned magnetic flux and magnetic energy existing in the photosphere at any given time \\citep[e.g.,][]{unn59,ste82,yi93,san98c,san04,sch03b}. The importance of these ubiquitous quiet Sun magnetic fields depends, to a large extent, on the magnetic field strengths characterizing them. For example, the magnetic flux and the magnetic energy scale as powers of the field strength, and the connectivity between photosphere and corona is probably provided by the magnetic concentrations with the largest field strengths \\citep[e.g.,][]{dom06}. Unfortunately, measuring quiet Sun magnetic field strengths is not a trivial task. All measurements are based on the residual polarization left when a magnetic field of complex topology is observed with finite angular resolution \\citep[e.g.,][]{emo01,san03,tru04}. Measuring implies assuming many properties of the unresolved complex magnetic field and, in doing so, the measurements become model dependent and prone to bias. Depending on the technique used for measuring, the real quiet Sun exhibits weak daG fields \\citep[e.g.,][]{ste82,fau93,bom05}, intermediate hG fields \\citep[e.g.,][]{lin99,kho02,lop06}, or strong kG fields \\citep[e.g.,][]{san00,lit02,dom03b}. Such discrepancies can be naturally understood if the true quiet Sun contains a continuous distribution of field strengths going all the way from zero to two kG. Different techniques are biased differently and, therefore, they tend to pick out a particular part of the distribution. This scenario is very much consistent with realistic numerical simulations of magneto-convection \\citep[][]{cat99a,stei06,vog05,vog07}. In order to provide a comprehensive observational description of the quiet Sun magnetic field strengths, one has to combine different methods carefully chosen to have complementary biases. Then the full distribution can be assambled taking the biases into account \\cite[][]{dom06}. In an effort to complement the existing magnetic field strength diagnostic techniques, \\citet{lop02} proposed the use of spectral lines whose Zeeman patterns are sensitive to hyperfine structure (HFS). The formalism to deal with the HFS of spectral lines in magnetic atmospheres was developed more than thirty years ago by \\citet{lan75}. According to such formalism, the polarization of the HFS lines vary with magnetic field strength very differently from the lines commonly used in solar magnetometry. This unusual behavior was invoked by \\citet{lop02} when proposing the use of HFS Mn~{\\sc i} lines as a diagnostic tool for magnetic field strengths. L\\'opez Ariste and coworkers have applied the idea to measure magnetic field strengths in quiet Sun regions \\citep{lop06,lop07,ase07}. Since the number of observables is limited, they minimize the statistical error of the measuremet by minimizing the number of free parameters to be tuned. Milne-Eddington atmospheres (ME) are used to synthesize the polarization of the lines. The magnetic field is assumed to be constant and, therefore, the measurements provide some kind of weighted average of the true field strengths existing in the resolution elements. As we pointed out above, the topology of the quiet sun magnetic field is complex, with a distribution of field strengths and polarities coexisting in a typical resolution element. Then the ill-defined average provided by the Mn~{\\sc i} lines is expected to be biased toward a particular range of field strengths, as it happens with the rest of the quiet Sun measurements (e.g., the NIR Fe~{\\sc i} lines overlook kG magnetic concentrations, \\citealt{san00}, \\citealt{soc03}; the traditional visible Fe~{\\sc i} lines exaggerate the contribution of kG fields, \\citealt{bel03}, \\citealt{mar06}; the Hanle scattering polarization signals are not sensitive to hG and kG, \\citealt{fau01}, \\citealt{san05}). The bias presented by the HFS Mn~{\\sc i} lines is so far unknown, and the existing and forthcoming measurements based on those lines will be fully appreciated only when the sources of systematic effects are properly acknowledged and quantified. In order to explore the magnitude and the sense of the expected effects, we undertake the synthesis of Mn~{\\sc i} lines in a number of realistic quiet Sun atmospheres with complex magnetic field distributions. The main trends are presented here. The paper is organized as follows. Section~\\ref{code} describes the software developed to carry out the syntheses. First, a ME code is needed to compare our syntheses with the results in the literature. Then, a plane parallel one-dimensional (1D) code allows us to explore the influence of realistic thermodynamic conditions on the polarization of lines with HFS. Finally, a MIcro-Structured Magnetic Atmosphere (MISMA) code provides additional realism to the modeling since it includes coupling between magnetic field strengths and thermodynamic conditions, magnetic fields varying along and across the line-of-sight (LOS), mixed polarities in the resolution element, etc. All these codes are based on the original FORTRAN routine written by \\citet{lan78}. The analysis is focused on the line most often used in observations, namely, \\mni{5538}. We discuss its intensity and circular polarization, since the linear polarization signals are very weak and they remain undetected so far. Single component and multi-component syntheses of this line are described in \\S~\\ref{scs} and \\S~\\ref{mcs}, respectively. An exploratory attempt to consider unresolved mixed polarities and velocity fields is carried out in \\S~\\ref{comp_obs}. The polarization of other Mn~{\\sc i} lines is also considered in \\S~\\ref{other}. The implications of these syntheses are discussed in \\S~\\ref{conclusions}. ", "conclusions": "The Zeeman pattern of the Mn~{\\sc i} lines depends on hyperfine structure (HFS), which confers them a sensitivity to hG magnetic field strengths different from the lines traditionally used in solar magnetometry. This peculiarity has been used to measure magnetic field strengths in quiet Sun regions (see \\S~\\ref{intro}). The methods applied so far assume the magnetic field strength to be constant in the resolution element, an approximation driven by feasibility rather than based on physical or observational arguments. Actually, it is not a good approximation since the magnetic fields of the quiet Sun are expected to vary on very small spatial scales, with field strengths spanning from zero to 2kG. Under these extreme conditions, all diagnostic techniques employed in quiet Sun magnetometry are strongly biased towards a particular range of field strengths, and the Mn~{\\sc i} signals are not expected to be the exception. Consequently, the diagnostic content of the Mn~{\\sc i} lines cannot be fully exploited unless their biases are properly understood. Such task is undertaken in the paper by exploring the response of Mn~{\\sc i} lines in a number of realistic quiet Sun scenarios. Three complementary LTE synthesis codes have been written and tested (ME, 1D and MISMA; \\S~\\ref{code}). They provide a number of relevant results, the first one being the ability to reproduce all observed Mn~{\\sc i} HFS patterns. We reproduce the observed unpolarized line profile of nine assorted lines corresponding to all HFS sensitivities (\\S~\\ref{other}). The study is focused on the line most often used in magnetometry, \\mni{5538}, however its behavior should be representative of the other lines. According to the weak magnetic field approximation, the Stokes $V$ signals scale with the longitudinal component of the magnetic field, i.e., the magnetic field strength times the cosine of the magnetic field inclination with respect to the LOS. We verify that the scaling on cosine holds very tightly. The scaling on the magnetic field strength, however, breaks down soon (when $B \\ge 400$~G for \\mni{5538}). Even for ME atmospheres, the weak field approximation predicts a Stokes~$V$ signal twice as large as the synthetic one for $B\\simeq$1.5~kG. When the expected coupling between the thermodynamic conditions and the magnetic field strengths is taken into account, the dimming of the kG Stokes~$V$ signals can be as large as two orders of magnitude (see Figs.~\\ref{maxv} and \\ref{phi}). The dimming of the polarization signals formed in kG magnetic concentrations affects all Mn~{\\sc i} lines (\\S~\\ref{scs}), and it has significant observational implications. If the resolution elements contains both weak and strong fields, then the kG fields tend to be under-represented in the average profile. We have modelled the bias assuming a multi component atmosphere, where the synthetic signals are weighted means of the Stokes profiles corresponding to each single field strength. The weight is given by the fraction of atmosphere filled by each field strength, i.e., by the magnetic field strength probability density function PDF (equations~[\\ref{pdfIV}] and [\\ref{pdfIVb}]). According to our modeling, even when the (unsigned) magnetic flux and the magnetic energy are dominated by kG magnetic fields, the Stokes~$V$ profile of \\mni{5538} can show HFS reversal at line core characteristic of hG fields (Fig.~\\ref{vsme}). A pure morphological inspection of the Stokes profiles does not suffice to infer which is the dominant magnetic field strength in the resolution element. For the HFS hump of \\mni{5538} to disappear the kG filling factor has to be larger than the sub-kG filling factor and, consequently, when the HFS hump disappears the magnetic flux and magnetic energy of the atmosphere are completely dominated by kG fields (\\S~\\ref{proxy}). Detecting Stokes~$V$ profiles with HFS features indicates the presence of hG fields in the resolution element. However, this sole observation does not tell whether the hG field strengths dominate. There seem to be two extreme alternatives to exploit the diagnostic potential of these lines. First, improving the spatial resolution of the observations to a point where the quiet Sun magnetic structures can be regarded as spatially resolved. Unfortunately, this possibility does not seem to feasible at present. Realistic simulations of magneto-convection indicate that quiet Sun magnetic fields are uniform only at the diffusive length scales \\citep[e.g.][]{cat99a,vog07}, which are of the order of a few km in the photosphere \\citep[e.g.][]{sch86}. These scales are very far from the angular resolution of the present measurements, and even much smaller than the length-scale for the radiative transfer average along the LOS \\citep{san96}. We prefer the alternative possibility, namely, developing inversion techniques where complex magnetic atmospheres are included into the diagnostics. Using appropriate constraints, the number of free parameters required to describe such atmospheres can be maintained within reasonable limits \\citep[e.g., the model MISMAs in ][]{san97b}. Dealing with unresolved velocities also favors detailled inversion codes (\\S~\\ref{comp_obs}). We are presently working on these improvements needed to develop the diagnostic technique pioneered by \\citeauthor{lop02}." }, "0710/0710.3326_arXiv.txt": { "abstract": "{I provide a short review of the properties of Narrow-line Seyfert 1 (NLS1) galaxies across the electromagnetic spectrum and of the models to explain them. Their continuum and emission-line properties manifest one extreme form of Seyfert activity. As such, NLS1 galaxies may hold important clues to the key parameters that drive nuclear activity. Their high accretion rates close to the Eddington rate provide new insight into accretion physics, their low black hole masses and perhaps young ages allow us to address issues of black hole growth, their strong optical FeII emission places strong constraints on FeII and perhaps metal formation models and physical conditions in these emission-line clouds, and their enhanced radio quiteness permits a fresh look at causes of radio loudness and the radio-loud radio-quiet bimodality in AGN. } \\resumen{Favor de proporcionar un resumen en espa\\~nol. \\\\ } \\addkeyword{Galaxies: Active} \\addkeyword{Galaxies: Seyfert} \\begin{document} ", "introduction": "\\label{sec:intro} Narrow-line Seyfert 1 galaxies are a subclass of active galactic nuclei (AGN). Their spectra exhibit exceptional emission-line and continuum properties. The most common NLS1 defining criterion is the width of the broad component of their optical Balmer emission lines in combination with the relative weakness of the [OIII]$\\lambda$5007 emission (FWHM$_{\\rm H\\beta} < 2000$ km/s and [OIII]/H$\\beta_{\\rm totl}$ $<$ 3; Osterbrock \\& Pogge 1985, Goodrich 1989){\\footnote{ While it is clear that a strict cutoff in line width (FWHM$_{\\rm H\\beta} < 2000$ km/s) is a gross simplification of any classification scheme, this historical value is still most commonly adopted for practical purposes. Suggestions have been made that more advanced NLS1 classification schemes would, for instance, incorporate the source luminosity (e.g., Laor 2000, Veron-Cetty et al. 2001). According to Sulentic et al. (2008 and references therein), AGN properties appear to change more significantly at a broad line width of FWHM$_{\\rm H\\beta} \\approx 4000$ km/s.}}. NLS1 galaxies typically show strong FeII emission which anticorrelates in strength with the [OIII] emission, and with the width of the broad Balmer lines. Often the presence of FeII emission is added as further NLS1 classification criterion and Veron et al. (2001) suggest the use of an intensity ratio FeII/H$\\beta_{\\rm totl} > 0.5$. NLS1 galaxies as AGN with the smallest Balmer lines from the Broad Line Region (BLR) and the strongest FeII emission, cluster at one extreme end of AGN correlation space. It is expected that such correlations provide some of the strongest constraints on, and new insights in, the physical conditions in the centers of AGN and the prime drivers of activity, and the study of NLS1 galaxies is therefore of particular interest. For instance, observations and interpretations hint at smaller black hole masses in NLS1 galaxies, and as such their black holes represent an important link with the elusive intermediate mass black holes, which have been little studied so far. Accreting likely at very close to the maximum allowed values, NLS1 galaxies are important test-beds of accretion models. This paper provides a short overview of the multi-wavelength properties of NLS1 galaxies and major models to explain them. ", "conclusions": "-- {\\sl{It seems hard to resist the feeling that nature is telling us something important here, but we do not yet know what it is.}} ~~~~~~~~~~~~~~Lawrence et al. (1997) \\vskip0.3cm While our knowledge has increased significantly in the last decade, important questions are still open. For instance, what are sufficient, what are necessary conditions for the onset of NLS1 activity ? For instance, the question is raised whether there are two types of Seyfert galaxies with low black hole masses: there is the unavoidable low-black-hole mass extension of BLS1 galaxies. Such systems would have their FWHM$_{\\rm H\\beta}$ fall below the formal cutoff value of 2000 km/s. Does low black hole mass already imply the emergence of some or all of the typical observed NLS1 characteristics ? Or is there a separate class of NLS1 galaxies ? While many individual NLS1 galaxies have been studied in great detail, we still need larger samples free of selection biases and well-suited BLS1 comparison samples in order to identify robust trends. Correlation space will ultimately have to be expanded to include the radio and infrared properties of NLS1 galaxies, as well as the properties of their host galaxies. On the theoretical/modeling side, interesting questions that persist or have emerged are related to mechanisms of super-Eddington accretion, the simultaneous presence of (TeV) blazar activity and high accretion rate in extreme radio loud NLS1 galaxies, and mechanisms of fueling and feedback in NLS1 galaxies. The study of NLS1 galaxies will continue to provide important contributions to our understanding of AGN and their cosmic evolution." }, "0710/0710.1609_arXiv.txt": { "abstract": "We examine the early phases of two near-limb filament destabilizations involved in coronal mass ejections on 16 June and 27 July 2005, using high-resolution, high-cadence observations made with the Transition Region and Coronal Explorer (TRACE), complemented by coronagraphic observations by Mauna Loa and the SOlar and Heliospheric Observatory (SOHO). The filaments' heights above the solar limb in their rapid-acceleration phases are best characterized by a height dependence $h(t)\\propto t^m$ with $m$ near, or slightly above, 3 for both events. Such profiles are incompatible with published results for breakout, MHD-instability, and catastrophe models. We show numerical simulations of the torus instability that approximate this height evolution in case a substantial initial velocity perturbation is applied to the developing instability. We argue that the sensitivity of magnetic instabilities to initial and boundary conditions requires higher fidelity modeling of all proposed mechanisms if observations of rise profiles are to be used to differentiate between them. The observations show no significant delays between the motions of the filament and of overlying loops: the filaments seem to move as part of the overall coronal field until several minutes after the onset of the rapid-acceleration phase. ", "introduction": "Observations of the early rise phase of filaments and their overlying fields can in principle help constrain the mechanisms involved in the destabilization of the magnetic configuration through comparison with numerical simulations (e.g., \\citep{fan2005}; \\citep{torok+kliem2005}; \\citep{williams+etal2005}; and references therein), because the detailed evolution depends sensitively on the model details. For example, a power-law rise with an exponent $m=2.5$ was obtained for a slender flux tube in the two-dimensional version of the catastrophe model (\\citep{Priest&Forbes02}). An MHD instability triggered by an infinitesimal perturbation implies an exponential rise, as was verified, for example, for a three-dimensional flux rope subject to a helical kink instability (\\citep{Torok&al04}, \\citep{torok+kliem2005}). The same holds for the torus (expansion) instability (TI), which starts as a $\\sinh(t)$ function (\\citep{Kliem&Torok06}) that is very similar to a pure exponential early on. The CME rise in a breakout model simulation was well described by a parabolic profile (\\citep{Lynch&al04}). The early rise phase of erupting filaments is best observed near the solar limb using high-resolution data, both in space and in time. Such data can be obtained by, for example, Big Bear Solar Observatory H$\\alpha$ observations (e.g., \\citep{kahler+etal1988}), the Mauna Loa K-coronameter (e.g., \\citep{gilbert+etal2000}), the Nobeyama Radioheliograph (e.g., \\citep{gopalswamy+etal2003}; \\citep{kundu+etal2004}), and the Transition Region and Coronal Explorer, TRACE (e.g., \\citep{vrsnak2001}; \\citep{gallagher+etal2003}; \\citep{goff+etal2005}; \\citep{sterling+moore2004}; \\citep{sterling+moore2005}; \\citep{williams+etal2005}). In those few cases where observers had the field of view for an appropriate diagnostic to attempt to establish whether the high loops or the filaments were accelerated first, the temporal resolution often was not adequate (see, e.g., \\citep{sterling+moore2004}, who use the standard 12-min.\\ cadence of SOHO/EIT). These studies show that filaments that are about to erupt often --~but not always~-- exhibit a slow initial rise during which both the filament and the overlying field expand with velocities in the range of $1-15$\\,km/s. Then follows a rapid-acceleration phase during which velocities increase to a range of $100$\\,km/s up to over $1000$\\,km/s. The rapid-acceleration phase finally transitions into a phase with a nearly constant velocity or even a deceleration into the heliosphere. The height evolution immediately following the onset of the rapid acceleration phase is often approximated by either an exponential curve (e.g., \\citep{gallagher+etal2003}; \\citep{goff+etal2005}; \\citep{williams+etal2005} --~who also show systematic deviations from that fit up to 2$\\sigma$ in position~-- ) or by a constant-acceleration curve (e.g., \\citep{kundu+etal2004}; and \\citep{gilbert+etal2000} --~ who show one case in which a third-order curve improves the fit to the earliest evolution, and leave others for future analysis); \\citet{kahler+etal1988} fit curves for the acceleration $a=ct^b$ to the first $10-50$\\,Mm for four erupting filaments, but do not list the best-fit values. \\cite{alexander&al02} find a best fit for the height of the early phase of a CME observed in X-rays by YOHKOH's SXT of the form $h_0+v_0t+ct^{3.7\\pm0.3}$. For 184 prominence events observed by the Nobeyama Radioheliograph, \\citet{gopalswamy+etal2003} show that higher in the corona velocity profiles include decelerating, constant velocity, and accelerating ones for heights from $\\sim 50$\\,Mm to 700\\,Mm above the solar surface. In many cases, the detailed study of the evolution of the early phase is hampered by insufficient temporal coverage or by gaps between the fields of view of two complementing instruments that can be as large as a few hundred Mm. This results in substantial uncertainties in the height evolution. \\citet{vrsnak2001}, for example, concludes that ``[t]he main acceleration phase \\ldots\\ is most often characterized by an exponential-like increase of the velocity'', but notes that polynomial or power-law functions fit at comparable confidence levels. In this study, we examine two events displaying the early destabilization and acceleration of ring filaments leading to coronal mass ejections. The high cadence down to 20\\,s, and the high spatial resolution of 1\\,arcsec, for the early evolution result in relatively small uncertainties in the height profiles. This enables a sensitive test of the height evolution against exponential, parabolic, and power-law fits. We find that a power-law with exponent near $3$, or slightly higher, is statistically preferred in both cases. As no published model matches that profile, we experiment with a numerical model for the torus instability, and find that this model can indeed approximate the observations provided that a sufficiently large initial velocity perturbation is applied (without which an exponential-like profile would be found). This finding reminds us of the sensitivity of developing instabilities to both initial and boundary conditions, and shows that the models, particularly their parametric dependencies, need to be worked out in greater detail in order to use observations of the height-time observations to differentiate successfully between competing models. ", "conclusions": "\\label{sec:conclusions} We study two well-observed filament eruptions, and find that their rapid acceleration phases are well fit by a cubic height-time curve that implies a nearly constant jerk for $10-15$\\,minutes, followed by a transition to a terminal velocity of $\\sim 750$\\,km/s and $\\sim 1250$\\,km/s, respectively. Simulations of a torus instability (TI) can reproduce such a behavior, provided that a substantial initial velocity perturbation is introduced. Without that perturbation, an exponential rise profile would be found. We note that the initial slow rise and the onset of the subsequent rapid acceleration phase are shared between the filament and overlying loop structures: neither \\referee{leads the other to within the temporal resolution. For characteristic Alfv{\\'e}n speeds over active regions of $\\sim 1,000$\\,km/s, the propagation of a perturbation over the separation of $\\sim 75,000$\\,km would require only $\\sim 1.2$\\,min., which corresponds to only one or two exposures. Thus the observations allow for Alfv{\\'e}nic propagation of a signal between filament and overlying loops, but suggest no longer-term differential evolution.} We observe no significant changes in the separation of erupting filament and overlying loops within that interval (Figs.~\\ref{fig:1c} and~\\ref{fig:2c}). After that, the distance increases in the 2005/06/16 eruption, suggesting the overlying field moves to the side for some time faster than the filament rises. For the 2007/07/27 eruption, the distance stays the same for one loop and decreases for another for up to 10~min after the start of the rapid acceleration phase, which reflects the significant sideways motion component of the rising filament. The observed configuration of the filament and high loops may be part of a larger overall destabilizing field configuration. Our numerical modeling has assumed that, in the rapid acceleration phase, the overlying field starts to move rapidly only as a consequence of the flux rope's destabilization. This is consistent with the data. However, we cannot exclude that the filament and the overlying field were destabilized simultaneously by a process different from the one considered here. More study is needed to establish whether the common evolution of the filament and high loops has a significant diagnostic value as to the cause of the instability. Comparison with other model studies in the literature leads us to conclude that the catastrophe model and the TI model are both marginally consistent with the observations of the two erupting filaments. The catastrophe model predicts a power-law exponent near the lower edge of the range of acceptable fits, but we have to allow for the possibility that changing that model's details may change the acceleration profile. \\referee{In order to yield the observed nearly cubic power-law rise (with $m$ slightly exceeding 3), our TI model requires an initial perturbation velocity that is in agreement with the observed rise velocity at the onset of the rapid-acceleration phase. If a nearly exact cubic rise were to be matched, however, initial velocities moderately exceeding the observed ones, by a factor $\\approx1.5$, were required. In any case, our modeling is consistent with the observed velocities after the first few minutes of the eruption.} Having established that the model for the TI instability is very sensitive to the initial conditions, we should of course also acknowledge that it depends sensitively on the model details itself. These include the details of the external field and of the rates and locations of the reconnection that occurs behind the erupting filament. That such reconnection occurs in reality is suggested for both events by the occurrence of brightenings mainly at the bottom side of the filaments at the onset of the rapid-acceleration phase. These brightenings develop later into the streaks used for position determination in Sect.~\\ref{sec:Observations}. The onset of reconnection even before the rapid-acceleration phase of the filament eruption on 27 July 2005 is strongly suggested by precursor soft and hard X-ray emission during about 04:00--04:30~UT, whose analysis revealed heating to 15~MK and the acceleration of non-thermal electrons to energies $>10$~keV (\\citep{Chifor&al06}). The observed rise velocity early in the filament eruption may be an underestimate of the true expansion velocity of the hoop formed by the flux rope: the filament channel in the pre-eruption phase of AR\\,10775 is strongly curved, and one of the two possible channels in AR\\,10792 is too (ambiguity exists here because the eruptions occurred very near the limb, so that the configurations of the filament channels can only be observed some days before and after the events, respectively). If the initial expansion of the flux rope would have a strong component in the general direction of the inclined plane of the curved filament channel rather than be purely normal to the solar surface, projection effects could cause us to underestimate the expansion velocity in particular early in the evolution. In addition to that, we must realize that the TI model assumes a flux rope that stands normal to the solar surface and that erupts radially. Future more detailed modeling will have to show how deviations from that affect the evolution of the eruption. The fact that the torus-instability model yields qualitatively different rise profiles (exponential vs.\\ power law) in different parts of parameter space, cautions against expectations that precise measurements of the rise profile of filament eruptions by themselves permit a determination of the driving process: the non-linearities in the eruption models clearly require high-fidelity modeling if such observations are to be used to differentiate successfully between competing models. Our initial modeling discussed here suggests that the torus instability is a viable candidate mechanism for at least some filament eruptions in coronal mass ejections. \\referee{Given the dependence of nonlinear models on the details of boundary and initial conditions, it will be necessary to investigate how other models for erupting filaments compare to the data, as well as how the fidelity of our modeling of the torus instability can be improved before we can reach definitive conclusions about the mechanism(s) responsible for filament eruptions in general.}" }, "0710/0710.4506_arXiv.txt": { "abstract": "The hydrogen-deficiency in extremely hot post-AGB stars of spectral class PG1159 is probably caused by a (very) late helium-shell flash or a AGB final thermal pulse that consumes the hydrogen envelope, exposing the usually-hidden intershell region. Thus, the photospheric elemental abundances of these stars allow to draw conclusions about details of nuclear burning and mixing processes in the precursor AGB stars. We compare predicted elemental abundances to those determined by quantitative spectral analyses performed with advanced non-LTE model atmospheres. A good qualitative and quantitative agreement is found for many species (He, C, N, O, Ne, F, Si, Ar) but discrepancies for others (P, S, Fe) point at shortcomings in stellar evolution models for AGB stars. PG1159 stars appear to be the direct progeny of [WC] stars. ", "introduction": "The PG1159 stars are a group of 40 extremely hot hydrogen-deficient post-AGB stars. Their effective temperatures ($T_{\\rm eff}$) range between 75\\,000 and 200\\,000~K. Many of them are still heating up along the constant-luminosity part of their post-AGB evolutionary path in the HR diagram ($L \\approx 10^4$L$_\\odot$) but most of them are already fading along the hot end of the white dwarf cooling sequence (with $L$ {\\raisebox{-0.4ex}{${\\stackrel{>}{\\scriptstyle \\sim}} $}} 10\\,L$_\\odot$). Luminosities and masses are inferred from spectroscopically determined $T_{\\rm eff}$ and surface gravity ($\\log g$) by comparison with theoretical evolutionary tracks. The position of analysed PG1159 stars in the ``observational HR diagram'', i.e., the $T_{\\rm eff}$--$g$ diagram, are displayed in Fig.\\,\\ref{fighrd}. The high-luminosity stars have low $\\log g$ ($\\approx$\\,5.5) while the low-luminosity stars have a high surface gravity ($\\approx$\\,7.5) that is typical for white dwarf (WD) stars. The derived mean mass is 0.57\\,M$_\\odot$, a value that is practically identical to the mean mass of WDs \\citep{be:07}. The PG1159 stars co-exist with hot central stars of planetary nebulae and the hottest hydrogen-rich (DA) white dwarfs in the same region of the HR diagram. About every other PG1159 star is surrounded by an old, extended planetary nebula. For a recent review with a detailed bibliography see \\citet{werner:06}. What is the characteristic feature that discerns PG1159 stars from ``usual'' hot central stars and hot WDs? Spectroscopically, it is the lack of hydrogen Balmer lines, pointing at a H-deficient surface chemistry. The proof of H-deficiency, however, is not easy: The stars are very hot, H is strongly ionized and the lack of Balmer lines could simply be an ionisation effect. In addition, every Balmer line is blended by a Pickering line of ionized helium. Hence, only detailed modeling of the spectra can give reliable results on the photospheric composition. The high effective temperatures require non-LTE modeling of the atmospheres. Such models for H-deficient compositions have only become available in the early 1990s after new numerical techniques have been developed and computers became capable enough. \\begin{figure*}[bth] \\vspace{-3.1cm} \\begin{center} \\epsfxsize=1.0\\textwidth \\epsffile{kwerner02.eps} \\end{center} \\vspace{-6.9cm} \\caption{\\label{fig:hrd} Complete stellar evolution track with an initial mass of 2\\,M$_\\odot$ from the main sequence through the RGB phase, the HB to the AGB phase, and finally through the post-AGB phase that includes the central stars of planetary nebulae to the final WD stage. The solid line represents the evolution of a H-normal post-AGB star. The dashed line shows a born-again evolution of the same mass, triggered by a very late thermal pulse, however, shifted by approximately $\\Delta \\log T_{\\rm eff} = -0.2$ and $\\Delta \\log~L/$L$_\\odot = - 0.5$ for clarity. The double-loop structure of the path is the consequence of a hydrogen-ingestion flash. The ``$\\star$'' symbol shows the position of PG1159$-$035 \\citep[from][]{werner:06}. } \\end{figure*} The first quantitative spectral analyses of optical spectra from PG1159 stars indeed confirmed their H-deficient nature \\citep{werner:91}. It could be shown that the main atmospheric constituents are C, He, and O. The typical abundance pattern is C=0.50, He=0.35, O=0.15 (mass fractions). It was speculated that these stars exhibit intershell matter on their surface, however, the C and O abundances were much higher than predicted from stellar evolution models. It was further speculated that the H-deficiency is caused by a late He-shell flash, suffered by the star during post-AGB evolution, laying bare the intershell layers. The re-ignition of He-shell burning brings the star back onto the AGB, giving rise to the designation ``born-again'' AGB star \\citep{iben:83a}. If this scenario is true, then the intershell abundances in the models have to be brought into agreement with observations. By introducing a more effective overshoot prescription for the He-shell flash convection during thermal pulses on the AGB, dredge-up of carbon and oxygen into the intershell can achieve this agreement \\citep{herwig:99c}. Another strong support for the born-again scenario was the detection of neon lines in optical spectra of some PG1159 stars \\citep{werner:94}. The abundance analysis revealed Ne=0.02, which is in good agreement with the Ne intershell abundance in the improved stellar models. If we do accept the hypothesis that PG1159 stars display former intershell matter on their surface, then we can in turn use these stars as a tool to investigate intershell abundances of other elements. Therefore these stars offer the unique possibility to directly see the outcome of nuclear reactions and mixing processes in the intershell of AGB stars. Usually the intershell is kept hidden below a thick H-rich stellar mantle and the only chance to obtain information about intershell processes is the occurrence of the third dredge-up. This indirect view onto intershell abundances makes the interpretation of the nuclear and mixing processes very difficult, because the abundances of the dredged-up elements may have been changed by additional burning and mixing processes in the H-envelope (e.g., hot-bottom burning). In addition, stars with an initial mass below 1.5~M$_\\odot$ do not experience a third dredge-up at all. The central stars of planetary nebulae of spectral type [WC] are believed to be immediate progenitors of PG1159 stars, representing the evolutionary phase between the early post-AGB and PG1159 stages. This is based on spectral analyses of [WC] stars which yield very similar abundance results (see papers by Crowther, Todt, and Gr\\\"afener in these proceedings). \\begin{figure*} \\begin{center} \\epsfxsize=1.0\\textwidth \\epsffile{kwerner03.eps} \\end{center} \\vspace{-0.5cm} \\caption{\\label{fig:psifn} Detail from FUSE spectra of two relatively cool PG1159 stars (see labels). Note the following features. The \\ion{F}{vi}~1139.5~\\AA\\ line which is the first detection of fluorine at all in a hot post-AGB star; the \\ion{P}{v} resonance doublet at 1118.0 and 1128.0~\\AA, the first discovery of phosphorus in PG1159 stars; the \\ion{N}{iv} multiplet at 1132~\\AA. Also detected are lines from \\ion{Si}{iv} and \\ion{S}{vi}. The broader features stem from \\ion{C}{iv} and \\ion{O}{vi} \\citep{reiff:07}.} \\end{figure*} ", "conclusions": "It has been realized that PG1159 stars exhibit intershell matter on their surface, which has probably been laid bare by a late final thermal pulse. This provides the unique opportunity to study directly the result of nucleosynthesis and mixing processes in AGB stars. Abundance determinations in PG1159 stars are in agreement with intershell abundances predicted by AGB star models for many elements (He, C, N, O, Ne, F, Si, Ar). For other elements, however, disagreement is found (Fe, P, S) that points at possible weaknesses in the evolutionary models. Generally, the abundance patterns clearly support the idea that [WC] stars are direct progenitors of PG1159 stars." }, "0710/0710.1786_arXiv.txt": { "abstract": "{A large fraction of the stellar mass is found to be located in groups of the size of the Local Group. Evolutionary status of poor groups is not yet clear and many groups could still be at an early dynamical stage or even still forming, especially the groups containing spiral and irregular galaxies only.} {We carry out a photometric study of a poor group of late-type galaxies around IC 65, with the aim: (a) to search for new dwarf members and to measure their photometric characteristics; (b) to search for possible effects of mutual interactions on the morphology and star-formation characteristics of luminous and faint group members; (c) to evaluate the evolutionary status of this particular group.} {We make use of our $BRI$ CCD observations, DPOSS blue and red frames, and the 2MASS $JHK$ frames. In addition, we use the \\ion{H}{i} imaging data, the far-infrared and radio data from the literature. Search for dwarf galaxies is made using the SExtractor software. Detailed surface photometry is performed with the MIDAS package. } {Four LSB galaxies were classified as probable dwarf members of the group and the $BRI$ physical and model parameters were derived for the first time for all true and probable group members. Newly found dIrr galaxies around the IC 65 contain a number of \\ion{H}{ii} regions, which show a range of ages and propagating star-formation. Mildly disturbed gaseous and/or stellar morphology is found in several group members. } {Various structural, dynamical, and star-forming characteristics let us conclude that the IC 65 group is a typical poor assembly of late-type galaxies at an early stage of its dynamical evolution with some evidence of intragroup (tidal) interactions.} ", "introduction": " ", "conclusions": "The main contributions of this paper are as follows: \\begin{enumerate} \\item We have selected four LSB dwarf companion candidates of the IC 65 group of galaxies on deep DPOSS frames according to their surface brightnesses, colours and morphology. \\item The $B, R$ and $I$ band surface photometry is presented for the first time for all certain and probable members of the group. The bright group members were studied in the NIR $J, H$ and $K$ bands, too. An image gallery and the deduced $SB$ and colour profiles are shown, permitting the detailed morphological analysis of the galaxies studied. Their relevant physical and model characteristics are determined. \\item Dynamical masses and star-forming characteristics of the bright group members are estimated using the new optical photometry and the available NIR, FIR, \\ion{H}{i} and radio data. The probable evolutionary status of the group is discussed. \\end{enumerate} An analysis of the available photometric and kinematic data of individual galaxies with emphasis to study the possible mutual interactions between the group members leads to the following results: \\noindent $\\bullet$ The available \\ion{H}{i} imaging data show that all bright members and at least one dwarf companion have a nearly normal gaseous fraction with \\ion{H}{i} mass to blue luminosity ratios in the range of 0.3 -- 1.0, consistent with their morphological type. The outer \\ion{H}{i} isophotes of the IC 65, and especially of the UGC 622 appear disturbed, in agreement with perturbation estimates.\\\\ $\\bullet$ The optical morphology of the bright galaxies generally appears to be regular, with barely significant disturbances in isophotes of the outer stellar disk of the IC 65 and UGC 622.\\\\ $\\bullet$ All bright group members (except PGC~138291, which we could not study in such detail) consist of many blue star-forming knots and plumes, especially UGC~608. A comparison of the surface photometry with stellar population models of Bruzual \\& Charlot (\\cite{bruzual03}) indicates that these blue knots must have formed recently. % The available data do not allow to establish whether they formed simultaneously, e.g. in star-bursts possibly triggered by interactions.\\\\ $\\bullet$ Two dIrr galaxies around the IC 65 both contain a number of \\ion{H}{ii} regions, which show a range of stellar ages and provide an evidence of propagating star-formation. One of these galaxies - A~0101+4744 is a confirmed member of the group; the second one - A~0100+4734 appears to be located in front of the group.\\\\ $\\bullet$ The brightest galaxies in both subgroups can fuel their current star-forming rates of $\\sim$ 1 - 2 ${\\cal M}_\\odot$ yr$^{-1}$ for about the next 3 - 7 Gyr. The IC~65 group of galaxies obviously belongs to the class of less evolved groups. It is composed of late type spiral and irregular galaxies arranged in two subgroups. No massive early type galaxies are present. No hot gas has been detected in it by the ROSAT survey. % Some morphological irregularities and signs of enhanced SF in its members could be indicative of recent/ongoing mutual interactions. Yet, the individual group members have retained much of their initial gas component. A few available velocities point to a short crossing time of only $\\sim$ 0.1~$H_0^{-1}$. However, this hardly means that the group has already reached a stable (virialized) configuration. The evidence, discussed above, lets us conclude that the IC~65 group of galaxies is a dynamically young system at a still relatively early stage of its collapse." }, "0710/0710.3783_arXiv.txt": { "abstract": "We report the first results of the GammeV experiment, a search for milli-eV mass particles with axion-like couplings to two photons. The search is performed using a ``light shining through a wall'' technique where incident photons oscillate into new weakly interacting particles that are able to pass through the wall and subsequently regenerate back into detectable photons. The oscillation baseline of the apparatus is variable, thus allowing probes of different values of particle mass. We find no excess of events above background and are able to constrain the two-photon couplings of possible new scalar (pseudoscalar) particles to be less than $3.1\\times 10^{-7} \\mbox{ GeV}^{-1}$ ($3.5\\times 10^{-7} \\mbox{ GeV}^{-1}$) in the limit of massless particles. ", "introduction": "The key to this experiment is the short 5 ns, 160 mJ pulses of 532 nm light emitted with a repetition rate of 20 Hz by our light source, a frequency-doubled Continuum Surelite I-20 Nd:YAG laser. As described below, the small duty cycle enables a large reduction in the detector noise via coincidence counting. The laser light is vertically polarized and when needed, a halfwave plate is used to obtain horizontal polarization. The laser pulses are sent through a vacuum system (diagrammed in Fig.~\\ref{F:apparatus}) designed around an insulating warm bore inserted into a 6 m Tevatron superconducting dipole magnet. The magnet produces a 5 T vertical field uniform across the aperture of the $48 \\mbox{ mm}$ inner diameter warm bore. A ``wall'' consisting of a high-power laser mirror on the end of a long (7 m) hollow stainless steel ``plunger'' is inserted into the warm bore. The mirror may be placed at various positions within the magnet by sliding the plunger. The plunger mirror projects the reflected wave into a photon state and the transmitted wave into a (pseudo-)scalar state, provided that the scalar is sufficiently weakly-interacting to pass through the material of the mirror. The mirror also functions to reflect the incident laser power out of the magnet to prevent heating of the magnet coils. The mirror is mounted on a welded stainless steel cap on the end of the plunger in order to prevent stray photons from passing through. Thus, the beam passing through the end of the plunger is a pure scalar beam. These scalars can then oscillate back into photons through the remaining magnetic field region inside the $35 \\mbox{ mm}$ inner diameter hollow plunger. Upon exiting the magnetic field region, the interaction ceases and the photon-scalar wavefunction is frozen. This combined wavefunction then propagates $\\sim 6\\mbox{ m}$ into a dark box where a Hamamatsu H7422P-40 photomultiplier tube (PMT) module is used to detect single photons in coincidence with the laser pulses. As described below, a high signal to noise ratio is achieved due to the very short pulses emitted by the laser. The photon-scalar transition probability may be written in convenient units as: \\begin{eqnarray} \\label{E:conversionprob1} P_{\\gamma \\rightarrow \\phi} &=& \\frac{4 B^2\\omega^2}{M^2 (\\Delta m^2)^2} \\sin^2 \\left( \\frac{\\Delta m^2 L}{4\\omega} \\right) \\\\ \\label{E:conversionprob2} &\\approx& \\frac{4 B^2\\omega^2}{M^2 m_\\phi^4} \\sin^2 \\left( \\frac{m_\\phi^2 L}{4\\omega} \\right) \\\\ \\label{E:conversionprob3} &=& 1.5\\times 10^{-11} \\frac{(B/\\mbox{Tesla})^2(\\omega/\\mbox{eV})^2}{(M/10^5 \\mbox{ GeV})^2 (m_\\phi/10^{-3} \\mbox{ eV})^4} \\nonumber \\\\ & & \\times \\sin^2 \\left( 1.267 \\frac{(m_\\phi/10^{-3} \\mbox{ eV})^2 (L/\\mbox{m})}{(\\omega/\\mbox{eV})} \\right) \\end{eqnarray} where $B$ is the strength of the external magnetic field, $\\omega$ is the inital photon energy, $L$ is the magnetic oscillation baseline. The mass-squared difference between the scalar mass and the effective photon mass, $\\Delta m^2=m_\\phi^2-m_\\gamma^2$, characterizes the mismatch of the phase velocities of the photon wave and the massive scalar wave and determines the characteristic oscillation length. While the photon does not really gain a mass in a normal dielectric medium, the phase advance may be modelled with an effective imaginary mass $m_\\gamma^2 = -2 \\omega^2 (n-1)$, where $n$ is the index of refraction \\cite{vanBibber:1988ge}. Both the warm bore and the interior of the plunger are pumped to moderate vacuum pressures of less than $10^{-4} \\mbox{ Torr}$, and a conservative estimate gives $\\sqrt{-m_\\gamma^2} < 10^{-4} \\mbox{ eV}$. Therefore, starting with Eqn.~\\ref{E:conversionprob2} we assume that the contribution from the effective photon mass is negligible for larger values of $m_\\phi$ near the PVLAS region. As can be seen from Eqn.~\\ref{E:conversionprob3}, the meter scale baseline provided by typical accelerator magnets is well-suited for probing the milli-eV range of possible particle masses. This fact can be a curse as well as a boon because for a monochromatic laser beam, a fixed magnet length may accidentally coincide with a minimum in the oscillation rather than a maximum. Indeed, this is a possible reason why the BFRT experiment \\cite{Ruoso:1992nx} did not see the PVLAS signal even though they had similar sensitivity. GammeV's plunger design allows us to change the oscillation baseline and thus scan through all possible values of the scalar mass in the milli-eV range without any regions of diminished sensitivity. The total conversion and regeneration probability contains two factors of Eqn.~\\ref{E:conversionprob2}, corresponding to the pre-mirror and post-mirror magnetic field regions of lengths $L_1$ and $L_2$. The total probability varies as $\\sin^2 \\left(\\frac {m_\\phi^2 L_1}{4\\omega} \\right) \\sin^2 \\left(\\frac{m_\\phi^2 L_2}{4\\omega} \\right)$ where $L_1+L_2=6\\mbox{ m}$. \\begin{figure}[t] \\includegraphics[width=0.48\\textwidth]{gammevcartoon.eps} \\caption{\\label{F:apparatus} Diagram (not to scale) of the experimental apparatus. The initial vacuum chamber consists of a 10 m insulating warm bore which is offset by 1.6 m from the end of the 6 m magnetic field region, and is sealed to the sliding plunger via a double o-ring assembly. The sliding plunger has a range of motion of 1.9 m, and contains an independent vacuum chamber. The vacuum window at the far end slides within a stationary, long dark box.} \\end{figure} To detect regenerated photons we use a $51\\mbox{ mm}$ diameter lens to focus the beam onto the $5 \\mbox{ mm}$ diameter GaAsP photocathode of the PMT. The alignment is performed using a low power green helium-neon alignment laser and a mock target. The alignment is verified both before and after each data-taking period by replacing the sealed plunger with an open-ended plunger, re-establishing the vacuum, and firing the Nd:YAG laser onto a flash paper target. An optical transport efficiency of $92\\%$ is measured using the ratio of laser power transmitted through the open-ended plunger and through the various optics and vacuum windows, to the initial laser power, using the same power meter in both cases to remove systematic effects. The quantum efficiency of the photocathode is factory-measured to be $38.7\\%$ while the collection efficiency of the metal package PMT is estimated to be $70\\%$. The PMT pulses are amplified by 46 dB and then sent into a NIM discriminator. Using a highly attenuated LED flasher as a single photon source, the discriminator threshold is optimized to give 99.4$\\%$ efficiency for triggering on single photo-electron pulses while also efficiently rejecting the lower amplitude noise. By studying the trigger time distribution, the deadtime fraction due to possible multiple rapid PMT pulses is found to be negligible (0.001$\\%$). Thus, we estimate the total photon transport and counting efficiency to be $(25\\pm 3)\\%$. Using this threshold, and the built-in cooler to cool the photocathode to $0^\\circ$C, we measure a typical dark count rate of 130 Hz. \\begin{figure}[t] \\includegraphics[width=0.48\\textwidth]{ttime_leaky_100ns.eps} \\caption{\\label{F:timing} PMT trigger times for the four run configurations, shown relative to the expected time distribution of photons as calibrated from the leaky mirror data.} \\end{figure} To perform the coincidence counting we use two Quarknet boards \\cite{quarknet} \\cite{quarknetgps} with 1.25 ns timing precision, referenced to a GPS clock. The Quarknet boards determine the absolute time of the leading edge of time-over-threshold triggers from the PMT and from a monitoring photodiode that is located inside the laser box. The clocks on the laser board and on the PMT board are synchronized using an external trigger from a signal generator. The absolute timing between the laser pulses and the PMT traces is established by removing the plunger with the mirror, and allowing the laser to shine on the PMT through several attenuation stages consisting of two partially reflective (``leaky'') mirrors, a pinhole, and multiple absorptive filters mounted directly on the aperture of the PMT module. The $10^{19}$ photons per second emitted by the laser are thus attenuated to a corresponding PMT trigger rate of less than 0.1 Hz for this timing calibration and to provide an {\\it in situ} test of the data acquisition system. The regenerated photons should arrive at the same time as the straight-through photons since milli-eV particles are also highly relativistic. For coincidence counting between the laser pulses and the PMT, a 10 ns wide window is chosen and includes $99\\%$ of the measured photon time distribution shown in Fig.~\\ref{F:timing}. The coincident dark count rate can be estimated to be $R_{\\rm{noise}}=20\\mbox{ Hz}\\times 130\\mbox{ Hz}\\times 10\\mbox{ ns} = 2.6\\times 10^{-5} \\mbox{ Hz}$. This noise rate is negligible to the expected signal rate of $\\sim 2\\times 10^{-3} \\mbox{ Hz}$ estimated from the central values of the PVLAS parameters. \\begin{table} \\begin{tabular}{|l|c|c|c|c|} \\hline Configuration&$\\#$ photons&Est.Bkgd&Candidates&g[GeV$^{-1}$]\\\\ \\hline Horiz.,center&$6.3\\times 10^{23}$&$1.6$&1&$3.4\\times 10^{-7}$\\\\ Horiz.,edge&$6.4\\times 10^{23}$&$1.7$&0&$4.0\\times 10^{-7}$\\\\ Vert.,center&$6.6\\times 10^{23}$&$1.6$&1&$3.3\\times 10^{-7}$\\\\ Vert.,edge&$7.1\\times 10^{23}$&$1.5$&2&$4.8\\times 10^{-7}$\\\\ \\hline \\end{tabular} \\caption{\\label{F:table} Summary of data in each of the 4 configurations.} \\end{table} ", "conclusions": "" }, "0710/0710.4450.txt": { "abstract": "In several studies of Low Luminosity Active Galactic Nuclei (LLAGNs), we have characterized the properties of the stellar populations in LINERs and LINER/HII Transition Objects (TOs). We have found a numerous class of galactic nuclei which stand out because of their conspicuous 0.1--1 Gyr populations. These nuclei were called ''Young-TOs'' since they all have TO-like emission line ratios. To advance our knowledge of the nature of the central source in LLAGNs and its relation with stellar clusters, we are carrying out several imaging projects with the Hubble Space Telescope (HST) at near-UV, optical and near-IR wavelengths. In this paper, we present the first results obtained with observations of the central regions of 57 LLAGNs imaged with the WFPC2 through any of the V (F555W, F547M, F614W) and I (F791W, F814W) filters that are available in the HST archive. The sample contains 34$\\%$ of the LINERs and 36$\\%$ of the TOs in the Palomar sample. The mean spatial resolution of these images is 10 pc. With these data we have built an atlas that includes structural maps for all the galaxies, useful to identify compact nuclear sources and, additionally, to characterize the circumnuclear environment of LLAGNs, determining the frequency of dust and its morphology. The main results obtained are: 1) We have not found any correlation between the presence of nuclear compact sources and emission-line type. Thus, nucleated LINERs are as frequent as nucleated TOs. 2) The nuclei of \"Young-TOs\" are brighter than the nuclei of \"Old-TOs\" and LINERs. These results confirm our previous results that Young-TOs are separated from other LLAGNs classes in terms of their central stellar population properties and brightness. 3) Circumnuclear dust is detected in 88$\\%$ of the LLAGNs, being almost ubiquitous in TOs. 4) The dust morphology is complex and varied, from nuclear spiral lanes to chaotic filaments and nuclear disk-like structures. Chaotic filaments are as frequent as dust spirals; but nuclear disks are mainly seen in LINERs. These results suggest an evolutionary sequence of the dust in LLAGNs, LINERs being the more evolved systems and Young-TOs the youngest. ", "introduction": "\\label{sec:Introduction} %{\\bf\\ojo\\ojon ROSA: YOUR FILE CAME WITH FUNNY LINE BREAKS, SO CHECK %THAT I HAVE NOT UNDONE SOME OF THE PARAGRAPH BREAKS INADVERTEDLY...} Low-luminosity active galactic nuclei (LLAGNs) comprise 30\\%\\ of all bright galaxies (B$\\leq$12.5) and are the most common type of AGN (Ho, Filippenko \\& Sargent 1997a, hereafter HFS97). These include LINERs, and transition-type objects (TOs, also called weak-[OI] LINERs). These two types of LLAGNs have similar emission line ratios in [OIII]/H$\\beta$, [NII]/H$\\alpha$, and [SII]/H$\\alpha$, but [OI]/H$\\alpha$ is lower in TOs than in LINERs. LLAGNs constitute a rather mixed class and different mechanisms have been proposed to explain the origin of the nuclear activity, including shocks, and photoionization by a non-stellar source, by hot stars or by intermediate age stars (e.g. Ferland \\& Netzer 1983; Filippenko \\& Terlevich 1992; Binnette et al. 1994; Taniguchi, Shioya \\& Murayama 2000). Because we do not know yet what powers them and how they are related to the Seyfert phenomenon, LLAGNs have been at the forefront of AGN research since they were first systematically studied by Heckman (1980). Are they all truly ``dwarf'' Seyfert nuclei powered by accretion onto nearly dormant supermassive black holes (BH), or can some of them be explained at least partly in terms of stellar processes? If LLAGNs were powered by a BH, they would represent the low end of the AGN luminosity function in the local universe and would also establish a lower limit to the fraction of galaxies containing massive BHs in their centers. If, on the contrary, LLAGNs were powered by nuclear stellar clusters, their presence would play an important role in the evolution of galaxy nuclei. Therefore, it is fundamental to unveil the nature of the central source in LLAGNs. There is clear evidence that at least some LINERs harbor a bona fide AGN, and they may be considered the faint end of the luminosity function of Seyfert galaxies. It has been found that about 20\\%\\ of the nearby LINERs have a weak broad H$\\alpha$ emission component similar to those found in type 1 Seyferts (Ho, Filippenko \\& Sargent 1997b). In a few of these LINERs the H$\\alpha$ line shows a double peak component (e.g. Storchi-Bergmann et al. 1997; Shields et al. 2000; Ho et al. 2000) suggesting that they are powered by an accreting black hole (BH). It has also been found that X-ray emission in LINERs has a nonthermal origin associated with an AGN (e.g. Terashima, Ho \\& Ptak 2000). Recent works based on higher spatial resolution data taken with Chandra show that only in half of the LLAGNs observed the X-ray emission is associated with compact nuclear cores (Satyapal et al. 2004; Dudik et al. 2006; Gonz\\'alez-Mart\\'\\i n et al. 2006). This is consistent with VLA radio observations, which detect unresolved radio cores in half of the LINERs (Nagar et al. 2000, 2002). Finally, a monitoring study at near-UV wavelengths by Maoz et al. (2005) finds UV variability in a significant fraction of the 17 LLAGNs observed. The variation of the UV fluxes may be interpreted as the manifestation of low rate or low radiative efficiency accretion onto a supermassive BH. On the other hand, detection of stellar wind absorption lines in the ultraviolet spectra of some TOs (Maoz et al. 1998; Colina et al. 2002) has proven unequivocally the presence of young stellar clusters in the nuclear region. Additional evidence comes from optical studies, in which we have focused on the study of the stellar population in the nuclei and circumnuclear region of LLAGNs to establish the role of stellar processes in their phenomenology. Ground-based (Cid Fernandes et al.\\ 2004; Cid Fernandes et al.\\ 2005) and HST+STIS spectra (Gonz\\'alez Delgado et al.\\ 2004) have shown that the contribution of an intermediate age stellar population is significant in a sizable fraction of the TO population. These studies identified a class of objects, called ``Young-TOs'', which are clearly separated from LINERs in terms of the properties and spatial distribution of the stellar populations. They have stronger stellar population gradients, a luminous intermediate age stellar population concentrated toward the nucleus ($\\sim$100~pc) and much larger amounts of extinction than LINERs. These objects, which underwent a powerful star formation event $\\sim$ 1 Gyr ago, could correspond to post-Starburst nuclei or to evolved counterparts of the Seyfert 2 with a composite nucleus, characterized too by harboring nuclear starbursts (Gonz\\'alez Delgado et al. 2001; Cid Fernandes et al. 2001, 2004). HST imaging of the nuclei of these Seyfert 2 galaxies shows that the UV emission is resolved into stellar clusters (Gonz\\'alez Delgado et al. 1998) that are similar to those detected in starburst galaxies (Meurer et al.\\ 1995). Nuclear stellar clusters are a common phenomenon in spirals, having been detected in 50-70\\% of these sources (Carollo et al. 1998, 2002; Boeker et al. 2002, 2004). Therefore stellar clusters are a natural consequence of the star formation processes in the central region of spirals. On the other hand, evidence has been accumulating during the past few years about the ubiquity of BH in the nuclei of galaxies. Furthermore, the tight correlation of the BH mass and stellar velocity dispersion (Ferrarese \\& Merrit 2000; Gebhardt et al.\\ 2000) implies that the creation and evolution of a BH is intimately connected to that of the galaxy bulge. Recently, in a HST survey in the Virgo Cluster, C\\^ot\\'e et al. (2006) have detected compact sources in a comparable fraction of elliptical galaxies. These compact stellar clusters, referred to as nuclei by the authors (see also Ferrarese et al 2006a), have masses that scale directly to the galaxy mass, in the same way as do the BH masses in high luminosity galaxies (Ferrarese et al. 2006b). Therefore, a natural consequence of the physical processes that formed present-day galaxies should be the creation of a compact massive object in the nucleus, either a BH and/or a massive stellar cluster. To determine the nature of the nuclear source of active galaxies and their evolution we are carrying out several projects with HST+ACS imaging at the near-UV wavelengths a sample of Seyferts (ID. 9379, PI. Schmitt, Mu\\~noz-Mar\\'\\i n et al. 2007) and LLAGNs (ID. 10548, PI. Gonz\\'alez Delgado). These observations are complemented with WFPC2 optical data retrieved from the HST archive. The high angular resolution provided by HST is crucial to determine the physical properties of the nuclei, the central structure of these galaxies, as well as to study the circumnuclear environment of AGNs. The main goals of these studies are to determine the frequency of nuclear and circumnuclear stellar clusters in AGNs, and whether they are more common in Seyferts, TOs or LINERs; to characterize the intrinsic properties of these clusters and to study whether there is evolution from Seyferts to TOs and LINERs. In addition, the frequency of dust and its morphology can also provide relevant information about the origin of nuclear activity. Dust is a valuable probe of the presence of cold interstellar gas in galaxies, and it is very sensitive to the perturbations that drive the gas toward the center and feed the AGN. Here, we present the initial results obtained for LLAGNs based on archival visible and red images obtained with the WFPC2. The paper is organized as follows: section 2 presents the sample selection, and 3 the characteristics of the observations. Sections 4, 5 and 6 describe the imaging atlas, the dust morphology and the central properties of the galaxies. Finally, the summary and conclusions are presented in section 7. %%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%%SEC%%% ", "conclusions": "LLAGNs, that include LINERs and TOs, are the most common type of AGN. What powers them is still at the forefront of AGN research. To unveil the nature of the central source we are constructing a panchromatic atlas of the inner regions of these galaxies, which will be used to determine their nuclear stellar population. To this end we have already carried out a near-UV snapshot survey of nearby LLAGNs with the ACS on board HST, that is complemented with optical and near-IR images available in the HST archive. In this paper, the first of a series, we present observations of 57 LLAGNs imaged with the WFPC2 through any of the V (F555W, F547M, F606W) and/or I (F791W, F814W) bands. These objects comprise 36$\\%$ of the original HFS97 LLAGNs sample, and correspond to those for which there are WFPC2 images available in the HST archive and whose circumnuclear stellar population we have already studied spectroscopically (Papers I--III). The subset of objects studied here follows the same distance and morphological type distribution of the complete HFS97 LLAGN sample. Classifying the objects in strong-[OI] and weak-[OI] ([OI]/H$\\alpha$)$\\leq$ 0.25), this subset includes 34$\\%$ and 36$\\%$ of the strong- and weak-[OI], respectively, of the whole HFS97 LLAGN sample. Following our results obtained from the analysis of the circumnuclear stellar population, this sub-sample contains 17 Young-TOs, 20 Old-TOs, 18 Old-LINERs and 2 Young-LINERs. Young-TOs or Young-LINERs are LLAGNs in which intermediate age stars contribute significantly to the nuclear blue-optical continuum. The dearth of Young-LINERs in the sample can be understood from the result obtained in our spectroscopic studies, which show that the overwhelming majority of LINERs harbor an old stellar population. With these data we have built an atlas that includes the structural map for all the images, and color maps for those galaxies for which images in two filters are available. We have identified those galaxies that have nuclear compact sources, and we have studied the circumnuclear environment of LLAGNs. We have found circumnuclear dust in 88$\\%$ of the LLAGNs, but this fraction is somewhat larger (95$\\%$) in weak-[OI] LLAGNs. The dust morphology is quite complex, from nuclear spiral lanes, to chaotic filaments and to nuclear disk-like structures. Chaotic filaments are as frequent as the well organized dust spirals; but disks are mainly seen in strong-[OI] LLAGNs. The dust concentration (simply graded by its location relative to a radius of 100-200 pc, is similar in weak- and in strong-[OI] because the fraction of LLAGNs with dust located only in the inner part is larger in Old-LLAGNs than in Young-LLAGNs. These results suggest an evolutionary dust sequence from Young-TOs to Old-LLAGNs. We have found that LINERs and TOs have both similar central magnitude and surface brightness, but LLAGNs with young and intermediate age populations are brighter than Old-TOs and LINERs. We have not found any correlation between the presence of nuclear compact sources and the emission line spectral type, ie., LINERs are as frequently nucleated as TOs. However, the centers of Young-TOs are brighter than the centers of Old-TOs and LINERs. The difference in magnitude and surface brightness can be even larger if we account for internal extinction, since Young-TOs are dustier. This result indicates that Young-TOs are separated from other type of LLAGNs also in terms of their central brightness, in addition of the properties and spatial distribution of the stellar population. These data have been very useful to study the circumnuclear environment of LLAGNs, and to identify which of these galaxies have a nuclear compact source. The fact that compact sources are as frequent in LINERs as in TOs, confirms again that LLAGNs are a mixed bag of objects. These results also suggest that the central morphology alone is not sufficient to elucidate the origin of their central source, and it cannot be used to ascertain whether LLAGNs are powered by AGNs or stellar clusters. These data will be complemented with near-UV (ACS) and near-IR (NICMOS) images to provide a panchromatic atlas of the inner regions of LLAGNs and to further investigate the origin of the nuclear sources and their relation with stellar clusters. {\\bf Acknowledgements} We thanks the referee for her/his suggestions that helped to improve the paper. RGD and EP acknowledge support from the Spanish Ministerio de Educaci\\'on y Ciencia through the grant AYA2004-02703. The data used in this work come from observations made with NASA/ESA Hubble Space Telescope, obtained from the STScI data archive. Basic research in radio astronomy at the NRL is supported by 6.1 base funding. We also thank support from a joint CNPq-CSIC bilateral collaboration grant. %%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%REF%%%" }, "0710/0710.4992_arXiv.txt": { "abstract": "We establish a nonminimal Einstein-Yang-Mills-Higgs model, which contains six coupling parameters. First three parameters relate to the nonminimal coupling of non-Abelian gauge field and gravity field, two parameters describe the so-called derivative nonminimal coupling of scalar multiplet with gravity field, and the sixth parameter introduces the standard coupling of scalar field with Ricci scalar. The formulated six-parameter nonminimal Einstein-Yang-Mills-Higgs model is applied to cosmology. We show that there exists a unique exact cosmological solution of the de Sitter type for a special choice of the coupling parameters. The nonminimally extended Yang-Mills and Higgs equations are satisfied for arbitrary gauge and scalar fields, when the coupling parameters are specifically related to the curvature constant of the isotropic spacetime. Basing on this special exact solution we discuss the problem of a hidden anisotropy of the Yang-Mills field, and give an explicit example, when the nonminimal coupling effectively screens the anisotropy induced by the Yang-Mills field and thus restores the isotropy of the model. ", "introduction": "The discussion of a nonminimal coupling (NMC) of gravity with fields and media has a long history. The most intensely this topic has been studied in connection with the problem of nonminimal coupling of gravity and scalar field, which has numerous cosmological applications. The details of the investigations of this problem are discussed, e.g., in the review of Faraoni {\\it et al\\/}.\\cite{FaraR} The development of the theory of NMC of gravity and scalar field $\\phi$ has started by the introduction of the term $\\xi \\phi^2 R$ to the Lagrangian ($R$ is the Ricci scalar). In Ref.~\\refcite{Chernikov} the special choice $\\xi = 1/6$ has been motivated by the conformal invariance; in Ref.~\\refcite{Callan} this quantity was considered as an arbitrary parameter of the model. Such a model has been widely used for the cosmological applications, in which $\\xi$ played a role of extra parameter of inflation (see, e.g., Refs.~\\refcite{Abbott}--\\refcite{Fara4}). In Refs.~\\refcite{HDehnen1}--\\refcite{HDehnen4} the gauge-invariant term $\\alpha {\\bf \\Phi}^{+}{\\bf \\Phi}R $ has been introduced instead of $\\xi \\phi^2 R$ in the context of non-Abelian gauge theory (${\\bf \\Phi}$~is a multiplet of scalar complex Higgs fields interacting with gravity and spinor matter.) Subsequent generalizations have been related to the replacement of $\\xi \\phi^2$ by the function $f({\\Phi}^2)$ (see, e.g., Refs.~\\refcite{Bergmann}--\\refcite{Steinh}), as well as, to the inserting of the terms of the type $F({\\Phi}^2, {\\cal R})$ both linear and nonlinear in the Ricci scalar, Ricci and Riemann tensors (see, e.g., Refs.~\\refcite{Linde}--\\refcite{Inagaki}). The idea of nonminimal derivative coupling introduced in Ref.~\\refcite{Amen3} and developed further in Refs.~\\refcite{Capo1,Capo2} has enriched the NMC modeling by the terms $\\phi_{,ij..}$. Nonminimal cosmological models based on the formalism of derivative coupling are the multi-parameter ones and have supplementary abilities for a fitting of observational data. Let us note that the NMC of gravity and scalar field leads to the modifications of both the Klein-Gordon and the Einstein equations, and such modifications are of interest for various inflation scenarios. Thus, the modeling of nonminimal interactions of scalar and gravitational fields is one of the well established and physically motivated branch of modern cosmology. Natural extension of the nonminimal theory from the models with scalar fields coupled to curvature to the models describing scalar fields interacting with gauge fields has the same sound motivation and can disclose new aspects of cosmological dynamics. The study of the nonminimal coupling of gravity with electromagnetic field has another motivation and another history. In 1971 Prasanna\\cite{Prasa1} introduced the invariant $R^{ikmn}F_{ik}F_{mn}$ ($R^{ikmn}$ is the Riemann tensor, $F_{ik}$ is the Maxwell tensor) as a possible element of a Lagrangian, and then in Ref.~\\refcite{Prasa2} obtained the corresponding nonminimal one-parameter modification of the Einstein-Maxwell equations. In 1979 Novello and Salim\\cite{Novello1} proposed to insert the gauge non-invariant terms $R A^k A_k$ and $R^{ik}A_i A_k$ in the Lagrangian ($A_k$ is an electromagnetic potential four-vector). A qualitatively new step has been made by Drummond and Hathrell in Ref.~\\refcite{Drum}, where the one-loop corrections to the quantum electrodynamics (QED) are obtained, which take into account the nonminimal coupling of gravity and electromagnetism. The Lagrangian of such a theory happens to contain three fundamental $U(1)$-gauge-invariant scalars $R^{ikmn}F_{ik}F_{mn}$, $R^{ik}g^{mn}F_{im}F_{kn}$ and $RF_{mn}F^{mn}$ with coefficients reciprocal to the square of the electron mass. This Lagrangian had no arbitrary parameters, but curvature induced modifications of the electrodynamic equations gave the impetus to wide discussions about the formal structure of the nonminimal Lagrangian, basic evolutionary equations, breaking the conformal invariance and the properties of the photons, coupled to curvature in different gravitational backgrounds (see, e.g., Refs.~\\refcite{Acci1}--\\refcite{Lafrance}). The last paper revived, as well, the interest to the paradigm: curvature coupling and equivalence principle, various aspects of which are now discussed (see, e.g., Refs.~\\refcite{Prasa3,Solanki}). The QED-motivation of the use of the generalized Maxwell equations can also be found in the papers of Kosteleck\\'y and colleagues.\\cite{Kost1,Kost2} The effect of birefringence induced by curvature, first discussed in Ref.~\\refcite{Drum}, and some of its consequences for the electrodynamic systems have been investigated in Refs.~\\refcite{Balakin1}--\\refcite{Balakin5} for the case of pp-wave background. The generalization of the idea of nonminimal interactions to the case of torsion coupled to the electromagnetic field has been made in Refs.~\\refcite{Hehl1,Hehl2} (see, also, Ref.~\\refcite{Hehl3} for a review on the problem). To summarize we stress that the study of electrodynamic systems nonminimally coupled to the gravity field poses a natural question about curvature induced variations of photon velocity in the cosmological background. Since the interpretation of observational data in cosmology depends essentially on the velocity of photon propagation during different cosmological epochs, the modeling of nonminimal electrodynamic phenomena seems to be well motivated and interesting from physical point of view. Concerning the nonminimal Einstein-Yang-Mills (EYM) theory, we can distinguish between two different ways to establish it. The first way is the direct nonminimal generalization of the Einstein-Yang-Mills (EYM) theory. In the framework of this approach Horndeski\\cite{Horn} and M\\\"uller-Hoissen\\cite{MH} obtained the nonminimal one-parameter EYM model from a dimensional reduction of the Gauss-Bonnet action. Now the Gauss-Bonnet models are of great interest in connection with the problem of dark energy (see, e.g., the Gauss-Bonnet model with nonminimal scalar field\\cite{OdinDE}). Thus, the non-Abelian multi-parameter extensions of nonminimal models are also well motivated, since they give a chance to explain the accelerated expansion of the Universe without addressing to exotic substance. We follow the alternative way, which is connected with a non-Abelian generalization of the nonminimal Einstein-Maxwell theory along the lines proposed by Drummond and Hathrell\\cite{Drum} for the linear electrodynamics. Based on the results of Ref.~\\refcite{BL05} a three-parameter gauge-invariant nonminimal EYM model linear in curvature is considered.\\cite{1BZ06}\\cdash\\cite{BDZ07} Our goal is to formulate a nonminimal Einstein-Yang-Mills-Higgs (EYMH) theory, and this process, of course, also admits different approaches. In fact, the nonminimal EYMH theory should accumulate the ideas and methods both from the nonminimally extended EYM theory and from the nonminimally extended scalar field theory. Initial attempt to develop nonminimal EYMH theory can be found, for instance, in Ref.~\\refcite{Bij}, where the scalar Higgs field is nonminimally coupled with gravity via $\\xi {\\Phi}^2 R$ term, and the Higgs field ${\\bf \\Phi}$ is included into the Lagrangian of the Yang-Mills field in a composition with a square of the Yang-Mills potential: ${\\Phi}^2 A_k^{(a)} A^k_{(a)}$. Such a theory is not gauge-invariant. In this paper we establish a new six-parameter nonminimal Einstein-Yang-Mills-Higgs model. First three coupling parameters, $q_1$, $q_2$ and $q_3$, describe a nonminimal interaction of Yang-Mills field and gravitational field. The fourth and fifth parameters, $q_4$ and $q_5$, describe the so-called gauge-invariant nonminimal derivative coupling of the Higgs field with gravity. Since the gauge-invariant derivative, $\\D_m {\\Phi}^{(a)}$, contains the potential of the Yang-Mills field, the corresponding nonminimal term is associated with ``triple'' interaction, namely, gravitational and scalar fields, gauge and scalar fields, and gauge and gravitational fields. The sixth parameter, $\\xi$, is the well-known coupling parameter nonminimally connecting gravitational and scalar fields via the term $\\xi R {\\Phi}^2$. Of course, this model is only one of a wide class of the nonminimal EYMH models. As for its motivation and possible physical applications, one can see that on the one hand, the interest to a six-parameter nonminimal EYMH model is based on the sound results obtained earlier in the framework of partial nonminimal models (Einstein-Maxwell, Einstein-Yang-Mills and scalar field theories), on the other hand, the six-parameter model under discussion shows new specific solutions of cosmological type, which can not appear in more simple models. The paper is organized as follows. In Sec. \\ref{Formalism} we formulate the nonminimal EYMH model, which contains six phenomenological coupling parameters, and establish the nonminimally extended Yang-Mills, Higgs and Einstein equations. In Sec. \\ref{IsModel} we apply the introduced master equations to the spacetime with constant curvature and obtain the specific relationships between coupling constants, which turn the extended equations for the gauge field and scalar field into identities. In Subsec. \\ref{dSsptime} we discuss the exact solutions to the nonminimal EYMH equations attributed to the isotropic cosmological model with Yang-Mills field, characterized by hidden anisotropy. ", "conclusions": "\\label{Discussion} \\noindent \\par 1. The main mathematical result of the presented paper is the establishing of a new self-consistent nonminimal system of master equations for the coupled Yang-Mills, Higgs and gravity fields from the gauge-invariant nonminimal Lagrangian (\\ref{1act}). The obtained mathematical model contains six arbitrary parameters, and, thus, admits a wide choice of special sub-models interesting for the applications to the nonminimal cosmology (isotropic and anisotropic) and nonminimal colored spherical symmetric objects. The applications require the phenomenological coupling constants $q_1$, $q_2,\\ \\dots,\\ q_5$ and $\\xi$ to be interpreted adequately. Following the idea, discussed in Ref.~\\refcite{HDehnen4}, we intend not to introduce ``new constants of Nature'', but to relate the phenomenological parameters with the constants well-known in the High Energy Particle Physics, on the one hand, and with the constants of cosmological origin, on the other hand. Indeed, in the specific cosmological model, established above, the sixth phenomenological parameter $\\xi$ is expressed in terms of the square of the effective mass of the Higgs bosons $\\mu$ and constant curvature $K$, $\\xi = \\frac{\\mu}{12K}$. Other parameters are expressed in terms of $K$ (see (\\ref{q})). Since in the de Sitter model the Hubble constant is $H=\\sqrt{K}$, one can say that $q_1$, $q_2,\\ \\dots,\\ q_5$ are connected with $H$. Analogously, one can consider the equality $H^2=K=\\frac{\\Lambda}{3}$ and thus, one can say that they are connected with the cosmological constant $\\Lambda$. In any case the parameters of nonminimal coupling $q_1$, $q_2,\\ \\dots,\\ q_5$ can be expressed in terms of cosmological parameters $K$, $H$ or $\\Lambda$, and define a specific radius of curvature coupling, $r_q \\equiv \\frac{1}{\\sqrt{K}}$ and the corresponding time parameter $t_q \\equiv r_q/c$. 2. The curvature coupling modifies the master equations for the Yang-Mills and Higgs fields. According to (\\ref{Heqs}) a new tensor ${H}^{ik}_{(a)}$ appears (see (\\ref{HikR})), which is an analog of the induction tensor in the Maxwell theory\\cite{Maugin}. This means that the curvature coupling of the non-Abelian gauge field with gravity acts as a sort of quasi-medium with a nonminimal susceptibility tensor ${\\cal R}^{ikmn}$ (see (\\ref{HikR})). As well, the curvature coupling modifies the master equations for the Higgs field, and the tensor $\\Re^{mn}$, according to (\\ref{Heq}), can be indicated as a simplest nonminimal susceptibility tensor for the Higgs field, and the vector ${\\Psi}^{m}_{(a)}$ (see (\\ref{21Heq})) can be defined as scalar induction. For the specific set of coupling constants (see (\\ref{AQU}), (\\ref{q})) the non-Abelian induction $H^{ik}_{(a)}$ and the scalar induction ${\\Psi}^{m}_{(a)}$ can turn into zero, despite the fact that the Yang-Mills field strength ${F}^{ik}_{(a)}$ and the Higgs field ${\\Phi}^{(a)}$ are non-vanishing. This means that, when (\\ref{AQU}) holds, the possibility exists to satisfy the nonminimally extended Yang-Mills and Higgs equations for arbitrary ${F}^{ik}_{(a)}$ and ${\\Phi}^{(a)}$. This possibility gives, in principle, a new option for modeling physical processes in Early Universe and shows very interesting analogy between this nonminimal model and resonance phenomena in plasma physics. Indeed, when we deal with plasma waves (for instance, with the longitudinal waves) one can see that electric induction $\\vec{D}$ is connected with the longitudinal electric field $\\vec{E}_{||}$ with the frequency $\\omega$ by the relation $\\vec{D}= \\varepsilon_{||}\\vec{E}_{||}$. Here $\\varepsilon_{||}$ is the longitudinal dielectric permittivity, the simplest expression for this quantity can be obtained in the limit of long waves and gives $\\varepsilon_{||}= 1-\\frac{\\Omega^2_{p}}{\\omega^2}$, where $\\Omega_{p}$ is the well-known plasma frequency. When $\\omega=\\Omega_{p}$, one obtains $\\vec{D}=0$ and electrodynamic equations are satisfied for arbitrary $\\vec{E}_{||}$. Analogous feature can be found in the nonminimal model described above (see Eq. (\\ref{1simpa})). Indeed, the quantity $K$ with the dimensionality of squared frequency ($c=1$) can be regarded as an analog of $\\Omega^2_{p}$, the quantity $2(6q_1+3q_2+q_3)$ can be indicated as $1/\\omega^2$, then the term $1-2K(6q_1+3q_2+q_3)$ plays a role of effective permittivity scalar $\\varepsilon_q$. When this effective permittivity scalar vanishes, i.e., when the constants of nonminimal coupling are connected with the constant curvature $K$ according to (\\ref{AQU}), we obtain the resonance case, for which the Yang-Mills and Higgs equations are satisfied identically for arbitrary strength field tensor $F^{ik}_{(a)}$ and Higgs multiplet $\\Phi^{(a)}$, the color induction $H^{ik}_{(a)}$ being equal to zero. 3. The vector potential of the Yang-Mills field $ A^{(a)}_i$ enters the master equations via the gauge covariant derivative $\\hat{D}_k$, thus, the gauge field generates an anisotropy in the spacetime. Such an anisotropy, in general case, breaks down the symmetry of the model and produces the isotropy violation. Nevertheless, as it was shown above, the nonminimal coupling can effectively screen the anisotropy and guarantee the symmetry conservation. In the framework of this model one can speak about hidden anisotropy of the Yang-Mills field, keeping in mind that non-Abelian gauge field enters the master equations for the gravity field in the isotropic combinations only." }, "0710/0710.3867_arXiv.txt": { "abstract": "High speed collisions, although current in clusters of galaxies, have long been neglected, as they are believed to cause little damages to galaxies, except when they are repeated, a process called ``harassment\". In fact, they are able to produce faint but extended gaseous tails. Such low-mass, starless, tidal debris may become detached and appear as free floating clouds in the very deep HI surveys that are currently being carried out. We show in this paper that these debris possess the same apparent properties as the so-called ``Dark Galaxies\", objects originally detected in HI, with no optical counterpart, and presumably dark matter dominated. We present a numerical model of the prototype of such Dark Galaxies -- VirgoHI21 --, that is able to reproduce its main characteristics: the one-sided tail linking it to the spiral galaxy NGC 4254, the absence of stars, and above all the reversal of the velocity gradient along the tail originally attributed to rotation motions caused by a massive dark matter halo and which we find to be consistent with simple streaming motions plus projection effects. According to our numerical simulations, this tidal debris was expelled 750~Myr ago during a fly-by at 1100~km~s$^{-1}$ of NGC~4254 by a massive companion which should now lie at a projected distance of about 400~kpc. A candidate for the intruder is discussed. The existence of galaxies that have never been able to form stars had already been challenged based on theoretical and observational grounds. Tidal collisions, in particular those occurring at high speed, provide a much more simple explanation for the origin of such putative Dark Galaxies. ", "introduction": "With the availability of unprecedented deep HI blind surveys, a population of apparently free floating HI clouds without any detected stellar counterpart has become apparent \\citep{meyer04,davies04,things,alfalfa1,alfalfa2}. It has been suggested that a fraction of them could be ``dark galaxies\", a putative family of objects that would consist of a baryonic disk rotating in a dark matter halo, but that would differ from normal galaxies by being free of stars, having all their baryons under the form of gas. They would thus be ``dark'' in the optical and most other wavelengths, but visible through their HI emission, contrary to pure ``dark matter'' halos. Such dark galaxies would be extreme cases of Low Surface Brightness Galaxies (LSBs), a class of objects that have a particularly faint stellar content compared to their gaseous and dynamical masses \\citep[e.g.,][]{carignanfreeman88}. The formation of low mass dark galaxies is actually predicted by $\\Lambda$-CDM models \\citep[e.g.,][]{vandenbosch03,tully05}. \\citet{taylor05} provided theoretical arguments against the existence of galaxies that would have remained indefinitely stable against star formation, unless they are of very low mass, at least a factor ten below than that of classical dwarf galaxies. If they exist, the dark galaxies are predicted to have a low dynamical mass and HI content. In the Local Group, some possibly rotating, high-velocity clouds were speculated to be dark galaxies \\citep{simon04,simon06}. Further away, \\citet{davies06} argued that most previous HI blind surveys were not sensitive enough to rule out the existence of dark galaxies. And indeed, while the HIPASS survey failed at detecting HI~clouds without optical counterparts \\citep{doyle05}, deeper recent HI observations, in particular with the Arecibo telescope, have revealed a number of dark galaxy candidates \\citep{alfalfa2}. Among these free-floating low-mass HI clouds, one object located in the outerskirts of the Virgo Cluster has attracted much attention and discussion: VirgoHI21 \\citep[][see Fig.~1]{davies04,minchin05}. Despite a HI mass of only $\\sim 10^8$~M$_\\sun$, this elongated gaseous structure, mapped with the Westerbork Synthesis Radio Telescope (WSRT) by \\citet[][hereafter M07]{minchin07}, exhibits a velocity gradient as large as 220~km~s$^{-1}$ (see Fig.~\\ref{fig:pv-vhi21}). {\\it Assuming} that the observed HI velocities trace rotation, the inferred dynamical mass would be as large as $\\sim 10^{11}$~M$_\\sun$. The object shows no optical counterpart, even on deep HST images (M07). With such extreme properties, VirgoHI21 has become the prototype for dark galaxies, although its high dynamical mass is atypical even in models predicting the existence of Dark Galaxies. If real, an object like VirgoHI21 could tidally disturb the galaxies in their neighborhood, as investigated by \\cite{karachentsev06}. Actually VirgoHI21 itself lies at about 150~kpc from the massive spiral galaxy NGC~4254 (M~99), to which it is connected by a faint HI~filament. This structure could in principle be a bridge linking the two galaxies and would then appear as a sign of a tidal interaction between them (M07). \\begin{figure*} \\centering \\includegraphics[width=\\textwidth,angle=0]{f1.pdf} \\caption{The system VirgoHI21/NGC 4254. {\\it Left} The distribution of the atomic hydrogen is superimposed in blue on a true color optical SDSS image of the field acquired through the WIKISKY.org project. The HI maps actually combine two data set: the observations obtained with the Arecibo telescope as part of the ALFALFA project \\citep[courtesy of B. Kent,][]{haynes07}, which are sensitive enough to show the whole extent of the gaseous tail; the observations obtained at WSRT \\citep[courtesy of R. Minchin,][]{minchin07} which are not as deep, but have a much better spatial resolution. The HI maps were smoothed and the WRST data have been deconvolved by the elongated telescope beam. {\\it Right} Color coded first moment, velocity, field of the HI tail as mapped by Arecibo. The observed velocity range in km~s$^{-1}$ is indicated to the right.} \\label{fig:vhi21} \\end{figure*} \\begin{figure} \\centering \\includegraphics[width=\\columnwidth,angle=0]{f2.pdf} \\caption{Observed Position (x-axis) Velocity (y-axis) diagram along the position angle 22 degrees corresponding to the main direction of the HI bridge. As for Figure~\\ref {fig:vhi21}, the WSRT (gray) and Arecibo (light blue) data set have been superimposed to show all available information.}\\label{fig:pv-vhi21} \\end{figure} However, starless isolated gas clouds, showing a large velocity spread, are not necessarily genuine dark galaxies. Ram pressure can strip gas away from spirals in the vicinity of clusters, a process that does not affect stars. Interaction with an external field, for instance that of another galaxy, can expulse large amounts of material from the disk in the form of gas-rich tidal tails and debris. In that vein, \\citet[][hereafter B05]{bekki05} have suggested that interactions between flying-by (i.e. interacting without merging) galaxies orbiting in a potential well similar as the one produced by the Virgo Cluster form tidal tails that after some time can resemble isolated gas clouds containing little stars. Furthermore tidal tails are the place of large streaming motions, as shown for instance by the observations of the merger prototype NGC~7252 by \\citet{hibbard94} and the models of \\citet{bournaud04} and B05. The remaining HI structures of galaxy collisions can thus not only appear as isolated HI clouds, but also exhibit large velocity gradients that mimic those expected for rotating disks within dark matter haloes. In these conditions, serious doubts have raised whether VirgoHI21 is really a dark galaxy and not simply the result of a tidal interaction or of an harassment process in the Virgo Cluster (B05). They have recently been reinforced by the publication by \\citet{haynes07} of a deep HI map of the field, obtained with the Arecibo Telescope as part of the ALFALFA survey \\citep{alfalfa1,alfalfa2}. It reveals that VirgoHI21 actually lies within an even larger HI structure that extends further to the North, in the opposite direction of NGC~4254. This further suggests that the thin and long HI feature is in fact a tidal tail emanating from the spiral and that VirgoHI21 is just a denser cloud within it and thus not a dark galaxy. On the basis of HI spectra obtained with the Effelsberg telescope and a numerical model including the cluster ram pressure, \\citet{vollmer05} suggested that NGC~4254 has indeed interacted recently with a companion. However this study mostly focussed on the internal properties of the spiral and in particular the formation of VirgoHI21 was not modeled. Nevertheless, several arguments against a tidal origin for the HI cloud have been raised and addressed in detail in M07. The main ones regard the absence of a suitable interacting companion, the nonexistence of a counter tail -- a feature that is generally present in tidal interactions --, the total lack of stars in the HI tail/bridge, and, above all, the remarkable 200~km~s$^{-1}$ velocity gradient associated to VirgoHI21, which seems to be reversed and amplified with respect to the large scale velocity field along the rest of the HI structure. We will show in this paper that most of these criticisms actually apply to low-velocity encounters and not to the high-velocity ones, which are common in the cluster environment. High-velocity collisions have been neglected so far because they seem to cause little disturbances to stellar disks unless they are very numerous and participate to an harassment process \\citep[e.g.,][]{moore96}. To further investigate the role of high speed collisions in the formation of tidal debris, especially the gaseous ones, we have carried out a series of numerical simulations. We illustrate their impact showing a numerical model reproducing the morphology and kinematics of VirgoHI21. The numerical simulations are presented in Section~2. In Section~3, we compare the properties of tidal tails formed in high- and low-velocity galaxy encounters. In Section~4, we present the model which best fits VirgoHI21. Its actual nature is discussed in Section~5. Our conclusions and the implications for the detection of dark galaxies are summarized in Section~6. ", "conclusions": "\\subsection{From tidal tails to fake dark galaxies} \\label{fakedark} As shown in Section~\\ref{sect:highvel}, tidal tails, especially those formed during high-velocity encounters, share many of the properties expected for dark galaxies: starless HI features, strong velocity gradients due in one case to streaming motions and in the other to the presence of a massive dark matter halo. If furthermore some gaseous condensations are present in the tail, they may resemble isolated dark galaxies: indeed the bridge to the parent galaxy can be very faint and hard or impossible to detect on moderately deep HI maps. Tidal tails do not necessarily have uniform profiles; denser regions can even lie in their outermost regions \\citep[e.g.,][]{DBM04}. This can be because large pre-existing clouds from the parent disk are moved into the tail where they can form local overdensities \\citep{elmegreen93}, or because some parts of an initially uniform tail condense under the effect of gravity \\citep{BH92}. That these regions will lie far from the progenitor disk is a natural consequence of the extended flat rotation curves of spirals \\citep{DBM04}. Depending on their location and mass, these denser parts of tails can even be self-gravitating, form stars and become finally independent objects with the mass of dwarf galaxies: the so-called Tidal Dwarf Galaxies \\citep[TDGs,][for a review]{duc07}. In such cases, the internal motions within the young gravitationally bound object, in particular its rotation, will induce an additional velocity gradient. On the other hand, the less massive condensations that have not reached the critical HI column density threshold to form stars will appear as detached HI clouds without any stellar counterpart. Many of the known free-floating HI clouds, which are considered as Dark Galaxy candidates, have been found in clusters. In this environment, high velocity encounters have a high probability to occur, due to the large velocity dispersion of the cluster galaxies. Thus intra-cluster low-mass HI clouds of tidal origin could then be common but more dedicated studies should check this. Of course, this mechanism applies only if the parent galaxy had before the collision an extended HI disk. This requires in particular that it has not already crossed the cluster core where ram pressure would have contributed to truncate its gaseous disk. Tidal material generally quickly falls back onto the parent spiral, but this can take more than 2~Gyr in the outer parts \\citep[e.g.,][]{BD06}. The cluster tidal field can even prevent tidal debris from falling back \\citep{Mihos04}. This leaves time for fake dark galaxies of tidal origin to be observed while the interloper galaxy can be far away: at 1000~km~s$^{-1}$, it can be at a projected distance up to 2~Mpc two billion years later. \\bigskip \\subsection{The nature of VirgoHI21} After having argued that dark galaxies may in general be mistaken with tidal features and thus be fake ones, we discuss more specifically the nature of VirgoHI21, presenting pro and con arguments for the different scenarios proposed so far for its origin. \\subsubsection{Tidal debris?} M07 argued that the VirgoHI21 + HI~bridge system cannot be a tidal tail from the spiral NGC~4254, because of the following reasons: \\begin{itemize} \\item Interacting galaxies generally have pair of tidal tails, and indeed the models in B05 have two-tailed morphologies, while NGC~4254 has no counter-tail. \\item The tidal HI clouds in B05 models are star-poor but not star-free, while HST optical observations show VirgoHI21 being completely dark. \\item The velocity gradient along the HI~bridge gets reversed around the VirgoHI21 cloud, which seems to rotate in a direction opposite to the rest of the HI bridge. This reversal is said by M07 not to be explained by interaction models as those of B05. \\item The velocity spread ($\\Delta V=200$~km~s$^{-1}$) of VirgoHI21 could only be explained by an encounter at a comparable velocity (according to M07), implying that the interloper should not be further away than a few arcminutes and would have been identified. Moreover, low relative velocities are rare in the Virgo Cluster. \\item High velocity encounters are more common in this environment but velocities as large as $\\sim$~1000~km~s$^{-1}$ are ``far too large to generate tidal features such as bridges and tails\" (M07) . \\end{itemize} If indeed VirgoHI21 is a genuine dark galaxy, as claimed by M07, the HI~bridge would then be a tail expulsed from the dark galaxy by an interaction with NGC~4254 or the cluster field. We note here that our numerical model of VirgoHI21 which suggests that the tidal debris rather emanate from NGC~4254 addresses each of these above-mentioned concerns: \\begin{itemize} \\item The interaction occurred about 750~Myr ago, so that the counter-tail from the spiral is not necessarily expected to be observed anymore. By that time the interloper might also be far away. \\item The HI cloud formed in our model during a high--velocity encounter is free of old stars pre-existing to the galaxy interaction, and is not dense enough to form new stars. \\item The velocity spread ($\\Delta V=200$~km~s$^{-1}$) of the HI structure does not imply that the galaxy interaction had a similar velocity. It is in fact accounted for by a high-velocity encounter. \\item The reversal of the velocity gradient along the tail is reproduced thanks to projection effects. \\end{itemize} Whereas the global kinematics of VirgoHI21 and its bridge can be simply explained by streaming motions along a tidal structure, in detail, the model and the observations show some differences. They may be noted when comparing the observed (Fig.~\\ref{fig:pv-vhi21}) and simulated (Fig.~\\ref{fig:posvel}) Position--Velocity diagrams along the tail; in particular, as put forward by M07, the velocity gradient towards VirgoHI21 is locally larger within the HI cloud, while in our model it is not more enhanced at this location than further away near the tip of the tail. This local difference is actually not a real concern for our scenario. First, part of the local amplification of the gradient can result from the self-gravity of the VirgoHI21 cloud itself, that could be somewhat denser than in our model. In particular, as recently shown by \\citet{B07}, tidal debris may contain a significant dark component that has not been included in the simulations. Alternatively, the velocity field can be disturbed by objects in the neighborhood, for instance by the nearby dwarf galaxy, SDSS J121804.26+144510.4, also known as object~C (see Fig.~\\ref{fig:vhi21} and M07). The nature of this dwarf is actually unclear (see Appendix). We suppose here that it is a pre-existing object, physically interacting with the system. Object~C has a visible HI mass of $2\\times 10^7$~M$_\\sun$. We included its possible influence in a simple model. We describe it as a dwarf spheroid with a Plummer profile, total mass of $4\\times 10^8$~M$_\\sun$ (to include the dark matter mass) and scale-length of 5~kpc. It is located 8~kpc East from VirgoHI21 on the sky plane, and we assume it lies 20~kpc below VirgoHI21 in the radial direction. Simulating the whole trajectory of this dwarf in the simulation from $t=0$ would introduce too many parameters. Since we just want to illustrate its possible local effect, we simply add it as a fixed mass in the late stage, linearly increasing its mass from $0$ at $t=600$~Myr to the final value at $t=700$. The velocity perturbation induced by object~C is shown on Figure~\\ref{fig:posvel}. The model qualitatively reproduces the local amplification of the velocity gradient at the position of VirgoHI21, and in particular the ``S-shape\" of the velocity profile visible in the Position-Velocity diagram (see Fig.~\\ref{fig:pv-vhi21}). Tuning the parameters of the simulation may help to further reproduce the exact kinematical feature. Obviously other objects, such as NGC~4262 -- the gas--rich galaxy to the East in Figure~\\ref{fig:vhi21} -- might have also crossed the trajectory of the tidal tail and slightly interacted with it. In any case, whatever the real explication is, the ``S-shape\" of the velocity profile of VirgoHI21 is not inconsistent with a tidal origin. \\bigskip A second critical issue is the identification of the interloper responsible for the collision. In our high-velocity scenario, the interloper now lies at a projected distance of 400~kpc to the WNW of NGC~4254. A massive spiral is in fact present near this position: NGC~4192 (M~98). This galaxy, which is seen close to edge-on, has a maximum rotation velocity corrected for inclination of 236~km~s$^{-1}$, significantly higher than that of NGC~4254 (193~km~s$^{-1}$, according to the HyperLeda database), making it compatible with the 1.5:1 mass ratio used in our simulation. In our model, the interloper is today approaching us with a radial velocity of 715~km~s$^{-1}$ w.r.t. the target spiral. In the real Universe, NGC~4192 is indeed approaching, but with a relative velocity of $\\sim 2000$~km~s$^{-1}$ with respect to NGC~4254, i.e. much larger than in our model. This difference could be explained by the tidal field of the cluster which is not taken into account in our model. Adding it would introduce too many additional free parameters since the radial position and tangential velocity of the studied objects with respect to the cluster are unknown. The cluster gravitational field can modify some details in the properties of tidal tails in the long term \\citep{Mihos04}, but not their fundamental characteristics. However the cluster tidal field can have a more dramatic effect on the relative orbit of the distant galaxy pair over the nearly 750~Myr period since the encounter. Indeed, a cluster with the mass typical of Virgo, $M_\\mathrm{C}=10^{15}$~M$_\\sun$, at a distance $R_\\mathrm{C}=1$~Mpc from the pair, and a typical separation between the two galaxies $d\\sim 300$~kpc (which is the average 3-D separation from the time of the encounter until today) would create a tidal acceleration \\begin{equation} a_\\mathrm{t} \\simeq \\frac{G M_\\mathrm{C} d} {{R_\\mathrm{C}} ^3} \\end{equation} and, over a timescale of $T=750$~Myr, a relative velocity impulsion of \\begin{equation} \\Delta V_\\mathrm{t} \\simeq \\frac{G M_\\mathrm{C} d} {{R_\\mathrm{C}} ^3} \\times T \\end{equation} which would come in addition to the relative velocity found in our model. The values above imply $\\Delta V \\simeq 1100$~km~s$^{-1}$. This additionnal velocity difference would account for the observed relative velocities of NGC~4192/4254, and goes in the right direction if NGC~4254 lies behind the Virgo Cluster center. NGC~4192 is thus a fully possible interloper, in spite of its large distance and relative velocity. This galaxy does not show a strongly disturbed HI disk on moderately deep HI maps \\citep{Chung05}. One reason could be unfavorable internal/orbital parameters, for instance a retrograde orbit. Alternatively, faint tidal debris may be present, but only visible on deep HI observations, similar to those that revealed the existence of VirgoHI21. However, we do not intent to claim that NGC 4192 is the only possible interloper. Other combination of orbits/projection/age can certainly reproduce the properties of VirogHI21, and a galaxy flying away at $\\sim 1000$~km~s$^{-1}$ during $\\sim 1$~Gyr can be at a projected distance of 1~Mpc today, possibly even at the center of the Virgo Cluster. The number of massive interloper candidates is then large, making hard to identify the real culprit. This is anyway not required for our demonstration that VirgoHI21 can be a tidal debris, since we have shown that possible interlopers do exist. \\begin{figure} \\centering \\includegraphics[width=\\columnwidth]{f6.pdf} \\caption{Position--Velocity diagram of the gas for Model 7 at $t=$750 Myr when it best reproduces the actual morphology and velocity field of the system VirgoHI21/NGC 4254. The arrow indicates the axis of the ``Position\" direction. The PV diagram of a model where ``Object C\" has been artificially added during the simulation is also shown. }\\label{fig:posvel} \\end{figure} \\subsubsection{A kinematically decoupled Tidal Dwarf Galaxy?} The presence of a strong velocity gradient in a tidal tail, if not due to streaming motions, may actually pinpoint the presence of a gravitationally bound object that need not be a pre-existing dark-matter dominated galaxy. Massive substructures in tidal tails often become kinematically decoupled, self-gravitating and form new stars, becoming rotating Tidal Dwarf Galaxies. This is the case for VCC~2062, a TDG candidate in Virgo \\citep{Duc07b}. VirgoHI21 appears as a gas condensation within a tidal tail; it is currently not a TDG since it is starless, but could be its gaseous progenitor. Whether stars will be formed in this structure later is questionable. If the surface column density has remained unusually very low during the first several hundreds of Myr after the formation of VirgoHI21, it is unlikely that star-formation is ignited later on. On the other hand, the dynamical collapse time of the cloud, $1/\\sqrt{G \\rho}$ \\citep{elmegreen02}, is as large as 400 -- 500 Myr for an initially resting system and an average volume mass density in our modeled cloud of $\\rho \\sim 10^{-3}$~M$_{\\sun}$~pc$^{-3}$ (this is also about the density of the real VirgoHI21 cloud assuming a vertical scale-height of $\\sim$ 300~pc). Comparing this time scale to the age of the cloud, about 500~Myr (it appears in the model at $t=200-300$~Myr), one may conclude that VirgoHI21 would barely have had the time to collapse and form stars even in the most favorable conditions and incidentally that the absence of stars today is not dependent on an arbitrary choice of the threshold. In other words, the system could still be contracting under the effect of its internal gravity today, and begin to form stars later-on {\\it if} its density comes to exceed the star formation threshold. However, the main argument against VirgoHI21 being {\\it yet} a TDG fully responsible for the observed velocity gradient is the large dynamical mass inferred from the rotation curve: in galaxies made out of collisional debris, the dynamical mass should be of the same order as the luminous one, even if the presence of dark baryons may cause some differences between them \\citep{B07}. In the case of VirgoHI21, the dynamical mass inferred from the velocity curve is more than a factor of 3000 greater than the luminous one, i.e. that of the HI component. Clearly, streaming motions provide a much more reasonable explanation for the kinematical feature observed near VirgoHI21, if indeed this object is of tidal origin. \\subsubsection{Harrasment by the cluster field?} In the group environment, \\cite{bekki05b} proposed a scenario in which the group tidal field is able to strip gas from HI-rich galaxies, explaining the presence of isolated intergalactic HI clouds. Following this idea, B05 presented a model in which the combined action of galaxy-galaxy interactions and the cluster tidal field produce debris with properties similar to Dark Galaxies. \\citet{haynes07} even suggested that VirgoHI21 and the whole HI structure would result from just the long-term harassment by the large-scale cluster potential. However, the tidal field exerted by a structure of mass $M$ and typical scale $R$ scales as $M/R^3$. The tidal field of the Virgo Cluster ($10^{15}$~M$_\\sun$, 1~Mpc) at the present distance of NGC~4254 is then more than ten times smaller than that of the interloper galaxy in our interaction model ($2\\times 10^{12}$~M$_\\sun$, 50~kpc). It is just unlikely that the cluster field can develop a tail as long as the galaxy interaction can do. The harassment process has a longer timescale than the galaxy pair interaction, but over long timescales the orientation changes, which hardly accounts for the single, thin and long tail around NGC~4254. This structure is more typical of a short and violent interaction like a close galaxy encounter than a weaker and longer process like the harassment by the global cluster field. \\subsubsection{Ram pressure stripping?} \\label{rampressure} Ram pressure exerted by the cluster hot gas may also expulse gas from the outer regions of spiral disks and create isolated HI clouds without any optical counterpart. Yet, this scenario suffers fundamental concerns, also pointed out by M07, in particular: \\begin{itemize} \\item structures known to result from ram-pressure stripping are rather short and thick as observed in Virgo \\citep{crowl05,lucero05,vollmer06,chung07} and suggested by hydrodynamical simulations \\citep[e.g.,][]{Roediger07}, while the HI bridge of VirgoHI21 is much thinner and longer. \\item the kinematics is not directly explained too, in particular the reversing velocity gradient in the HI~bridge, which in the context of a tidal interaction results from streaming motions along a curved tail (in 3-D) more or less seen edge-on. This may also be the case for ram pressure but should be demonstrated by a model. \\end{itemize} \\cite{vollmer05} put forward the role of ram pressure, combined with a tidal interaction, in shaping the internal gas distribution, with its $m=1$ structure, and velocity field of NGC~4254. This partly explains why, in the innermost regions, the detailed kinematics of the spiral presents some differences with that of our model which did not take into account the intracluster medium. More recently, \\cite{Kantharia07} presented a low radio frequency continuum map of the galaxy which is best explained invoking a ram pressure scenario. However so far, its possible contribution on the properties of the HI bridge and VirgoHI21 has not yet been investigated. The two above-mentioned papers do not claim that their origin is ram-pressure, and indeed our pure tidal model is able to reproduce these features provided that the HI disk was originally much more extended than the optical radius -- an hypothesis which was not adopted in \\cite{vollmer05}. \\subsubsection{A dark galaxy?} Showing that virgoHI21 {\\it can} be a tidal debris taking the appearance of a dark galaxy does not directly rule out the possibility that it is a real dark galaxy. The dark galaxy hypothesis however suffers several difficulties unexplained so far: \\begin{itemize} \\item the maximal velocity gradient is not centered on the peak of the HI emission, but actually lies on one side of the VirgoHI21 cloud. This is unexpected for a rotating HI disk within a massive dark halo. One may propose that the on-going interaction with NGC~4254 causes this asymmetry in the velocity field, but that NGC~4254 is massive and close enough to induce such major disturbances remains to be shown. \\item within the assumption that VirgoHI21 is a real dark galaxy, the HI bridge is a tidal tail expulsed from it and captured by the more massive galaxy NGC~4254 (M07). The gas in the HI bridge, falling onto NGC~4254 from 150~kpc away, is not expected to have the same velocity as the local gas settled in rotation: it could be on retrograde, polar or direct orbits but with different velocities. However, the base of the HI bridge has a radial velocity coherent with the outer disk of NGC~4254 to which it is morphologically connected: the velocity step between the base of the tidal tail and the outer disk is in fact less than $50$~km~$^{-1}$, much smaller than the circular velocity there \\citep[see data in M07 and also][]{phookun93}. This further suggests that the HI material in the bridge comes from NGC~4254. \\end{itemize} These facts are naturally explained by the tidal scenario proposed for VirgoHI21. Whether they can also be addressed with the Dark Galaxy hypothesis remains to be demonstrated, in particular with a numerical model. Although challenged, the putative existence of Dark Galaxies as massive as VirgoHI21, has fostered a number of follow-up works. As discussed by \\citet{karachentsev06} such invisible ghost objects should tidally perturb galaxies in their neighborhood, explaining why a fraction of apparently isolated spiral stellar disks seem to show signs of an external perturbation. We note however that other mechanisms may account for them, such as accretion of diffuse gas \\citep{bournaud05m1}." }, "0710/0710.0675_arXiv.txt": { "abstract": "We investigate the number and type of pulsars that will be discovered with the low-frequency radio telescope LOFAR. We consider different search strategies for the Galaxy, for globular clusters and for galaxies other than our own. We show an all-sky Galactic survey can be optimally carried out by {\\em incoherently} combining the LOFAR stations. In a 60-day all-sky Galactic survey LOFAR can find over a thousand pulsars, probing the local pulsar population to a very deep luminosity limit. For targets of smaller angular size, globular clusters and galaxies, the LOFAR stations can be combined coherently, making use of the full sensitivity. Searches of nearby northern-sky globular clusters can find large numbers of low luminosity millisecond pulsars (eg.\\ over 10 new millisecond pulsars in a 10-hour observation of M15). If the pulsar population in nearby galaxies is similar to that of the Milky Way, a 10-hour observation can find the 10 brightest pulsars in M33, or pulsars in other galaxies out to a distance of 1.2Mpc. ", "introduction": "Since the discovery of the first four pulsars with the Cambridge radio telescope, an ongoing evolution of telescope systems has doubled the number of known radio pulsars roughly every 4 years. The next step in radio telescope evolution will be the use of large numbers of low-cost receivers that are combined to form an interferometer or a single dish. These telescopes, LOFAR \\citep{ls07}, the Allen Telescope Array \\citep{bow07} and the SKA \\citep{kram04}, create new possibilities for pulsar research. Here we outline and compare strategies for targeting normal and millisecond pulsars, both in the disk and globular clusters of our Galaxy, and in other galaxies. ", "conclusions": "Because of its large area and wide beam on the sky, LOFAR probes the local population of pulsars to a very deep luminosity limit. A 60-day Galactic survey at 140MHz can find over a thousand new pulsars, disclosing the local low-luminosity population. If we add all antennae coherently the sensitivity increases even further; with this setup, millisecond pulsars in nearby globular clusters can be detected to much lower flux limits than previously possible. Assuming the pulsar population in other galaxies is similar to that in ours, we can detect periodicities or giant pulses from extragalactic pulsars up to several Mpcs away." }, "0710/0710.3730_arXiv.txt": { "abstract": "Models of terrestrial planet formation in the presence of a migrating giant planet have challenged the notion that hot-Jupiter systems lack terrestrial planets. We briefly review this issue and suggest that hot-Jupiter systems should be prime targets for future observational missions designed to detect Earth-sized and potentially habitable worlds. ", "introduction": "Since the discovery of the first extrasolar planets \\citep{wolszczan,mayor}, astronomical techniques and observational baselines have advanced to the point where over 200 extrasolar planetary systems have been identified \\citep{butler}. Most detected exoplanets are in the giant planet mass range and it is now clear that our solar system is but one variant within a great diversity of planetary system architectures. One of the most surprising discoveries has been of a population of giant planets, the so-called \\emph{hot-Jupiters}, found orbiting in a region of extreme insolation very close ($r < 0.1~\\mathrm{AU}$) to their central stars and well within the radius of the original nebular snowline ($r \\approx 3 - 5~\\mathrm{AU}$) where giant planets are thought to form \\citep{pollack}. Hot-Jupiters are not uncommon: they amount to about a quarter of exoplanet discoveries, and are thought to provide evidence that protoplanets can migrate over large radial distances via tidal interactions with the protoplanetary disk \\citep[e.g.][]{lin1,lin2,ward2,nelson1}. Since the disk gas is observed to disperse within the first few Myr of the system's existence \\citep{haisch}, giant planets must form and migrate through the inner system within this period, which is considerably less than the $\\sim$~10--100~Myr thought to be required to complete terrestrial planet formation \\citep{chambers2,kleine,halliday,obrien}. Test particle studies have shown that terrestrial planets external to a hot-Jupiter would have stable orbits \\citep{jones}, and \\citet{raymond1} have shown that they should be able to form, in the presence of a giant planet already at $\\sim 0.1$~AU, from any available pre-planetary material with a period ratio roughly $>$~3. However, until recently it has been a common assumption that terrestrial planets could not have formed in hot-Jupiter systems due to the disruptive effect of the giant planet's migration which is deemed to have cleared the inner system of planet-forming material \\citep[e.g.][]{armitage1}, prompting claims that the observed abundance of hot-Jupiters could be used to constrain the general abundance of habitable planets \\citep{ward1}, and even their galactic location \\citep{lineweaver1}. This picture has been challenged by the work of two groups who have modeled terrestrial planet formation concurrently with, and following, an episode of giant planet migration \\citep{fogg1,fogg2,fogg3,fogg4,raymond2,mandell}. Their findings suggest that inner system solids disks are \\emph{not} destroyed by the intrusion of a migrating giant planet and that terrestrial planet formation can resume in the aftermath and run to completion. In this paper, we briefly describe our model of terrestrial planet formation and show some typical results. ", "conclusions": "Our models predict that terrestrial planets might be routinely expected in hot-Jupiter systems, including within their habitable zones, and may be detectable by forthcoming missions such as Kepler, Darwin, SIM PlanetQuest and TPF." }, "0710/0710.1503_arXiv.txt": { "abstract": "{Observations of water lines are a sensitive probe of the geometry, dynamics and chemical structure of dense molecular gas. The launch of Herschel with on board HIFI and PACS allow to probe the behaviour of multiple water lines with unprecedented sensitivity and resolution.} {We investigate the diagnostic value of specific water transitions in high-mass star-forming regions. As a test case, we apply our models to the AFGL\\,2591 region.} {A multi-zone escape probability method is used in two dimensions to calculate the radiative transfer. Similarities and differences of constant and jump abundance models are displayed, as well as when an outflow is incorporated.} {In general, for models with a constant water abundance, the ground state lines, i.e., $\\mathrm{1_{10}}$-$\\mathrm{1_{01}}$, $\\mathrm{1_{11}}$-$\\mathrm{0_{00}}$, and $\\mathrm{2_{12}}$-$\\mathrm{1_{01}}$, are predicted in absorption, all the others in emission. This behaviour changes for models with a water abundance jump profile in that the line profiles for jumps by a factor of $\\sim$\\,10-100 are similar to the line shapes in the constant abundance models, whereas larger jumps lead to emission profiles. Asymmetric line profiles are found for models with a cavity outflow and depend on the inclination angle. Models with an outflow cavity are favoured to reproduce the SWAS observations of the $\\mathrm{1_{10}}$-$\\mathrm{1_{01}}$ ground-state transition. PACS spectra will tell us about the geometry of these regions, both through the continuum and through the lines.} { It is found that the low-lying transitions of water are sensitive to outflow features, and represent the excitation conditions in the outer regions. High-lying transitions are more sensitive to the adopted density and temperature distribution which probe the inner excitation conditions. The Herschel mission will thus be very helpful to constrain the physical and chemical structure of high-mass star-forming regions such as AFGL\\,2591.} ", "introduction": " ", "conclusions": "We have constructed models to examine the excitation of water in the high-mass star-forming region AFGL\\,2591. Depending on the adopted density, temperature and abundance structure, a completely different set of line profiles and strengths is found. Hence, the line profiles are very sensitive to the adopted physical and chemical structure. We have found that ({\\it i}) the ground-state transitions 1$_\\mathrm{10}$-1$_\\mathrm{01}$, 2$_\\mathrm{12}$-1$_\\mathrm{01}$ and 1$_\\mathrm{11}$-0$_\\mathrm{00}$, with relatively low upper energy levels ($\\lesssim$\\,110\\,K), become highly optically thick in the outer regions. The line profiles for these transitions, are mainly dominated by the emission from the outer regions, and are therefore not useful to put constraints on the water abundance in the inner regions. However, ({\\it ii}) the emission from lines with higher upper energy levels is dominated by the emission originating in the inner regions, and are therefore useful to probe the water abundance in the warm inner regions. ({\\it iii}) For models with an outflow cavity, the outflow feature (blue peak less strong than the red peak) is best seen in the ground-state transitions of o- and p-H$_\\mathrm{2}$O. ({\\it iv}) The influence of a moderate disk (few 100 AU in size) in the centre of the AFGL\\,2591 region does not change the water line profiles and strengths within the Herschel beam. The Herschel mission will thus greatly help to understand the structure of high-mass protostellar objects, and consequently the formation process of high-mass stars." }, "0710/0710.1359_arXiv.txt": { "abstract": "In a systematic study, we compare the density statistics in high-resolution numerical experiments of supersonic isothermal turbulence, driven by the usually adopted solenoidal (divergence-free) forcing and by compressive (curl-free) forcing. We find that for the same rms Mach number, compressive forcing produces much stronger density enhancements and larger voids compared to solenoidal forcing. Consequently, the Fourier spectra of density fluctuations are significantly steeper. This result is confirmed using the $\\Delta$-variance analysis, which yields power-law exponents $\\beta\\!\\sim\\!3.4$ for compressive forcing and $\\beta\\!\\sim\\!2.8$ for solenoidal forcing. We obtain fractal dimension estimates from the density spectra and $\\Delta$-variance scaling, and by using the box counting, mass size and perimeter area methods applied to the volumetric data, projections and slices of our turbulent density fields. Our results suggest that compressive forcing yields fractal dimensions significantly smaller compared to solenoidal forcing. However, the actual values depend sensitively on the adopted method, with the most reliable estimates based on the $\\Delta$-variance, or equivalently, on Fourier spectra. Using these methods, we obtain $D\\!\\sim\\!2.3$ for compressive and $D\\!\\sim\\!2.6$ for solenoidal forcing, which is within the range of fractal dimension estimates inferred from observations ($D\\!\\sim\\!2.0\\dots2.7$). The velocity dispersion to size relations for both solenoidal and compressive forcings obtained from velocity spectra follow a power law with exponents in the range $0.4\\dots0.5$, in good agreement with previous studies. ", "introduction": "Observations provide velocity dispersion to size relations for various molecular clouds (MCs), which document the existence of supersonic random motions on scales larger than $\\sim\\!0.1\\,\\mathrm{pc}$ \\citep[e.g.,][]{Larson1981,Myers1983,PeraultFalgaronePuget1986,SolomonEtAl1987,FalgaronePugetPerault1992,HeyerBrunt2004}. These motions are associated with compressible turbulence \\citep[e.g.,][]{ElmegreenScalo2004,ScaloElmegreen2004,MacLowKlessen2004} in the interstellar medium \\citep{Ferriere2001} and exhibit a single turbulent cascade or spatially separated coexisting inertial ranges \\citep{PassotPouquetWoodward1988} similar to the kinetic energy cascade of incompressible \\citet{Kolmogorov1941c} turbulence. However, there are various physical processes (e.g., self-gravity, magnetic fields, nonequilibrium chemistry) and especially the compressibility of the gas, that alter the scaling laws \\citep[e.g.,][]{Fleck1996} and statistics \\citep[e.g., intermittency corrections measured by][]{HilyBlantFalgaronePety2008} established for incompressible turbulence. The physical origin and characteristics of the turbulent fluctuations are still a matter of debate. To advance on the question of how turbulence isdriven in the interstellar medium, we present results of high-resolution numerical experiments of supersonic isothermal turbulence comparing two distinct and extreme ways of driving the turbulence in a systematic study: 1) solenoidal forcing (divergence-free or rotational forcing), and 2) compressive forcing (curl-free or dilatational forcing). Various numerical and analytical studies have provided important insight into the statistics of supersonic isothermal turbulence \\citep[e.g.,][]{PorterPouquetWoodward1992,Vazquez1994,PadoanNordlundJones1997,PassotVazquez1998,StoneOstrikerGammie1998,MacLow1999,Klessen2000,OstrikerStoneGammie2001,BoldyrevNordlundPadoan2002,LiKlessenMacLow2003,PadoanJimenezNordlundBoldyrev2004,JappsenEtAl2005,BallesterosEtAl2006,KritsukEtAl2007,LemasterStone2008}. Most of these studies use purely solenoidal or weakly compressive kinetic energy injection mechanisms (forcing) to excite turbulent motions. In the present study, we aim at comparing the usual case of solenoidal (divergence-free) forcing with the case of fully compressive (curl-free) forcing. The actual way of turbulence production in real MCs is expected to be far more complex compared to what we can model with the present simulations, probably consisting of a convolution of various agents producing turbulence, and mixtures of solenoidal and compressive modes \\citep[e.g.,][]{ElmegreenScalo2004,MacLowKlessen2004}. Here, we systematically investigate the extreme cases of purely solenoidal versus purely compressive energy injection. Analyzing the density correlation statistics and fractal structure obtained in our hydrodynamic simulations, we show that compressive forcing leads to significantly steeper density fluctuation spectra and consequently to fractal dimensions of the turbulent gas structures, that are significantly smaller compared to the usually adopted solenoidal forcing. We use Fourier analysis, $\\Delta$-variance analysis, structure functions, the fractal mass size, box counting, and perimeter area methods to obtain fractal dimension estimates. We apply the $\\Delta$-variance analysis to both our 3-dimensional data and to 2-dimensional projections, and the perimeter area method to projections and slices through the turbulent density structures supporting the result of a significantly smaller fractal dimension for compressive forcing compared to solenoidal forcing. Although compressive forcing yields significantly smaller fractal dimensions than solenoidal forcing, the estimated fractal dimensions are in the range $2.0\\dots2.7$ consistent with observational estimates \\citep[e.g.,][]{ElmegreenFalgarone1996,SanchezEtAl2007} We explain our numerical method, construction of solenoidal and compressive forcing fields and fractal analysis techniques in Section~\\ref{sec:methods}. In Section~\\ref{sec:results}, we show that our results are consistent with previous studies using solenoidal forcing, whereas compressive forcing yields much stronger density contrasts and consequently leads to significantly smaller fractal dimensions. In Section~\\ref{sec:conclusions}, we summarize our conclusions. ", "conclusions": "\\label{sec:conclusions} We have presented results of two high-resolution ($1024^3$ grid cells) hydrodynamic simulations of supersonic isothermal turbulence driven to rms Mach numbers $\\mathcal{M}\\!\\sim\\!5.5$. The first simulation uses the typically adopted solenoidal (divergence-free) forcing to excite turbulent motions, whereas the second one uses compressive (curl-free) forcing. We have shown that compressive forcing yields much stronger density contrasts compared to solenoidal forcing for the same rms Mach number. This implies that the turbulence production mechanism leaves a strong imprint on compressible turbulence statistics, especially altering the density statistics. Our results particularly suggests that the mixture of solenoidal and compressive modes of the turbulence forcing must be taken into account. We summarize our results as follows: \\begin{itemize} \\item The velocity Fourier spectra exhibit power laws in the inertial range for solenoidal and compressive forcing. The slopes obtained for both forcings are significantly steeper ($\\sim\\!1.9$) compared to the Kolmogorov slope ($5/3$), in agreement with previous studies \\citep[e.g.,][]{KritsukEtAl2007,SchmidtEtAl2008} and in agreement with velocity dispersion to size relations inferred from observations \\citep[e.g.,][]{Larson1981,FalgaronePugetPerault1992,HeyerBrunt2004,PadoanEtAl2006}. \\item From the integral of the density fluctuation Fourier spectra and from the asymptotic behavior of the 2nd order density structure function, we obtained the standard deviation of the density distribution $\\sigma_\\rho$. Compressive forcing yields a standard deviation $\\sim\\!$ three times larger compared to solenoidal forcing, in agreement with the results found in our previous study analyzing density probability distribution functions \\citep{FederrathKlessenSchmidt2008} and in agreement with the studies by \\citet{PassotVazquez1998}, \\citet{KritsukEtAl2007}, \\citet{BeetzEtAl2008} and \\citet{SchmidtEtAl2008}. \\item The density fluctuation Fourier spectra are significantly steeper for compressive forcing in the inertial range compared to solenoidal forcing. Consistent results were obtained using complementary analysis methods, i.e., by comparing the $\\Delta$-variances \\citep{OssenkopfKripsStutzki2008a} and the 2nd order structure functions of the density field. Our estimates of density spectra for solenoidal forcing are in agreement with previous studies, e.g., the weakly magnetized super-Alfv\\'enic supersonic MHD models by \\citet{PadoanEtAl2004} and \\citet{KowalLazarianBeresnyak2007}, and consistent with the hydrodynamic estimates by \\citet{KritsukNormanPadoan2006}. Although a comparison with observational results must be regarded with caution due to systematic uncertainties, our results for solenoidal and compressive forcing are in the range of inferred scaling exponents by observations \\citep[e.g.,][]{BenschStutzkiOssenkopf2001}. \\item From the scaling of the density fluctuation Fourier spectra and the $\\Delta$-variance applied to the 3-dimensional data and applied to 2-dimensional projections, we obtained fractal Hurst exponents following the analysis by \\citet{StutzkiEtAl1998}. This implies fractal box counting and fractal perimeter area dimensions significantly smaller for compressive forcing compared to solenoidal forcing (see Table~1). \\item We analyzed the density structure using the fractal mass size method as introduced by \\citet{KritsukEtAl2007}. Compressive forcing yields a smaller fractal mass dimension compared to solenoidal forcing. The mass size method is, however, particularly sensitive to the temporal fluctuations of density peaks. Given the large uncertainties, our results using this method are roughly consistent with the estimates by \\citet{KritsukEtAl2007} and \\citet{KowalLazarian2007}. \\item We analyzed the fractal density structure using the box counting method described in Section~\\ref{sec:box-counting} and the perimeter area method (Section~\\ref{sec:perimeter-area}) applied to projections and slices. We recover the significant differences between solenoidal and compressive forcing inferred from the density spectra and $\\Delta$-variance analysis. However, the box counting dimension varies strongly with the defining density threshold. The perimeter area dimensions obtained from slices are roughly consistent with the computed perimeter area dimensions from the $\\Delta$-variance given the systematic uncertainties (of order $\\sim\\!0.1$ for fractal dimension estimates) comparing different methods. The range of fractal dimensions obtained is consistent with the observations analyzed by \\citet{ElmegreenFalgarone1996} suggesting an overall fractal dimension of interstellar clouds in the range $D\\!\\sim\\!2.3\\pm0.3$. \\end{itemize}" }, "0710/0710.3454_arXiv.txt": { "abstract": "{There are now four low mass X-ray binaries with black holes which show twin resonant-like HFQPOs. Similar QPOs might have been found in Sgr A$\\sp *$. I review the power spectral density distributions of the three X-ray flares and the six NIR flares published for Sgr A$\\sp *$ so far, in order to look for more similarities than just the frequencies between the microquasar black holes and Sgr A$\\sp *$. The three X-ray flares of Sgr A$\\sp *$ are re-analysed in an identical way and white noise probabilities from their power density distributions are given for the periods reported around $\\sim$ 1100 s. Progress of the resonant theory using the anomalous orbital velocity effect is summarized. ", "introduction": "% \\label{sect:intro} Quasi-periodic oscillations (QPOs) are believed to arise from variations of the accretion flow around compact objects, i.e. white dwarfs, neutron stars and black holes. As far as black holes are concerned, Remillard and McClintock (2006) have compiled a list of a total of 40 sources which show up in galactic X-ray binaries, 20 of which show a dark companion with a mass largely exceeding the mass of a neutron star but apparently limited to about 18 solar masses. Out of the 20 sources with established dynamically measured masses 16 sources show low frequency QPOs, whereas high frequency QPOs (HFQPOs, $\\nu$ $>$ 10 Hz) have been detected in 8 sources. ", "conclusions": "" }, "0710/0710.4274_arXiv.txt": { "abstract": "We performed MHD simulations of very light bipolar jets with density contrasts down to $10^{-4}$ in axisymmetry, which were injected into a medium of constant density and evolved up to $200$ kpc ($200\\: r_{\\mathrm{j}}$) full length. These jets show weak and roundish bow shocks as well as broad cocoons and thermalize their kinetic energy very efficiently. We argue that very light jets are necessary to match low-frequency radio observations of radio lobes as well as the bow shocks seen in X-rays. Due to the slow propagation, the backflows and their turbulent interaction in the midplane are important for a realistic global appearance. ", "introduction": "During the last years, simulations of extragalactic jets with reasonable resolution and realistic sizes became computationally feasible, which makes comparisons between simulated and observed properties possible \\citep{Saxton2002,Zanni2003,Carvalho2005,KrauseVLJ2,ONeill2005}. Unfortunately, the direct physical variables and the observed properties are rather hard to link, which leaves simulations with a wide range of parameters. Simulations are mainly governed by the initial setup of the density ratio between jet and ambient gas, the Mach number and the magnetic field. If the magnetic field is not dynamically dominant (though important), the density contrast is the most dominant parameter, but may be one of the hardest to measure. The thermal jet pressure has turned out to be of little importance in the very light jet limit \\citep{KrauseVLJ1}. As the (kinetic) power of a jet can be estimated from energies in X-ray bubbles, typical values of velocity, lifetime, jet radius and cluster gas densities indicate that density contrasts of $10^{-2}$ to $10^{-4}$ (or even lower) are necessary to describe real sources. Parameter studies support this further, if the global jet/cocoon/bow shock properties are compared. Thus, we concentrate on the very light jets with magnetic fields as another important ingredient. ", "conclusions": "" }, "0710/0710.3381_arXiv.txt": { "abstract": "We present optical and X-ray data for the first object showing strong evidence for being a black hole in a globular cluster. We show the initial X-ray light curve and X-ray spectrum which led to the discovery that this is an extremely bright, highly variable source, and thus must be a black hole. We present the optical spectrum which unambiguously identifies the optical counterpart as a globular cluster, and which shows a strong, broad [O III] emission line, most likely coming from an outflow driven by the accreting source. ", "introduction": "Since the early days of X-ray astronomy, there has been considerable debate over whether globular clusters contained black holes. With the discovery of Type I X-ray bursts from all globular clusters in the Milky Way with bright X-ray sources (starting with Grindlay et al. 1976), and their subsequent explanation as episodes of thermonuclear burning on the surfaces of neutron stars (Woosley \\& Taam 1976; Swank et al. 1977), it became clear that there was no evidence for any accreting black holes in the Milky Way's globular cluster system. Intepretations of the observations have been taken in two directions. One is simply that given only 13 bright X-ray sources in the Milky Way's globular cluster system, it is not so unlikely for them all to have neutron star accretors, especially in light of the fact that about 10 times as many neutron stars as black holes are expected to be produced for most stellar initial mass functions. The alternative is that dynamical effects are responsible for ejecting black holes from globular clusters. Severe mass segregation is likely to take place for globular cluster black holes, as they should be many times heavier than all the other stars in the cluster. This can lead to the formation of a ``cluster within a cluster'' where the heaviest stars (i.e. the black holes) feel negligble effects from the other stars in the cluster, which in turn leads to a cluster with a short evaporation timescale (Spitzer 1969). Numerical calculations have found that this evaporation can be accelerated further due to binary processes (e.g Portegies Zwart \\& McMillan 2000). Early results from the Chandra X-ray Observatory gave new hope that globular cluster black holes might be detectable, by opening up the window of looking in other galaxies. Previously, only ROSAT could resolve point sources in other galaxies, and its localization of sources was generally not good enough to allow for unique identification of optical counterparts. The first few years of Chandra observations revealed several extragalactic globular cluster X-ray sources brighter than the Eddington limit for a neutron star (e.g. Angelini et al. 2001; Kundu et al. 2002), but a globular cluster may contain multiple bright neutron stars (as does, for example M~15 in our own galaxy -- White \\& Angelini 2001), and that the quality of X-ray spectra available from Chandra for even the brightest extragalactic sources is insufficient to make phenomenological determinations that a source has a black hole accretor. It was pointed out that only large amplitude variability could prove that we were seeing the emission from a single source, rather than multiple sources (Kalogera, King \\& Rasio 2003). Furthermore, the optical catalogs used to identify globular cluster counterparts to X-ray sources have been predominantly photometric catalogs, sometimes made even without color selections being used to ensure that that the optical source in question truly is a globular cluster. Most studies done to date have focused on HST images of the central regions of elliptical galaxies with high specific frequencies of globular clusters. In these regions, and with the angular resolution of HST, contamination will be rare, especially if color cuts are used to ensure that the contribution of background quasars is minimized. In the halos of galaxies, the surface density of real globular clusters will drop, and contamination will be a more serious problem. In either case, when one is looking for conclusive proof that an object is a globular cluster black hole, spectroscopic confirmation that the object is a globular cluster is essential -- the fractional contamination of the X-ray sources due to background AGN will be more serious at very high fluxes, corresponding to luminosities above $10^{39}$ ergs/sec than it will at lower levels consistent bright neutron star accretors. Furthermore, the optical to X-ray ratios for globular cluster black holes near the Eddington limit and background quasars are quite similar. ", "conclusions": "We have shown for the first time clear evidence of a globular cluster black hole, on the basis of strong, highly variable X-ray emission from a source in a spectroscopically confirmed globular cluster. Based on the X-ray spectrum, the characteristic variability, and the [O III] emission in the optical spectrum, the black hole is most likely a stellar mass object accreting far faster than its Eddington rate. Given that it is difficult to develop a scenario in which a globular cluster could have both a stellar mass black hole in a binary and an intermediate mass black hole, this argues against the idea that all globular clusters contain intermediate mass black holes. This discovery also motivates future searches for quiescent stellar mass black holes in the Milky Way's globular clusters; these may be hiding among the X-ray sources currently classified as cataclysmic variable stars. Radio emission should be detectable only from quiescent stellar mass black holes, and should be the one feasible discriminant between the two classes of systems." }, "0710/0710.4935_arXiv.txt": { "abstract": "Intercluster filaments negligibly contribute to the weak lensing signal in general relativity (GR), $\\gamma_{N}\\sim 10^{-4}-10^{-3}$. In the context of relativistic modified Newtonian dynamics (MOND) introduced by Bekenstein, however, a single filament inclined by $\\approx 45^\\circ$ from the line of sight can cause substantial distortion of background sources pointing towards the filament's axis ($\\kappa=\\gamma=(1-A^{-1})/2\\sim 0.01$); this is rigorous for infinitely long uniform filaments, but also qualitatively true for short filaments ($\\sim 30$Mpc), and even in regions where the projected matter density of the filament is equal to zero. Since galaxies and galaxy clusters are generally embedded in filaments or are projected on such structures, this contribution complicates the interpretation of the weak lensing shear map in the context of MOND. While our analysis is of mainly theoretical interest providing order-of-magnitude estimates only, it seems safe to conclude that when modeling systems with anomalous weak lensing signals, e.g. the ``bullet cluster\" of Clowe et al., the ``cosmic train wreck\" of Abell 520 from Mahdavi et al., and the ``dark clusters\" of Erben et al., {\\it filamentary structures might contribute} in a significant and likely complex fashion. On the other hand, {\\it our predictions of a (conceptual) difference in the weak lensing signal could, in principle, be used to falsify MOND/TeVeS} and its variations. ", "introduction": "\\label{intro} Without resorting to cold dark matter (CDM), the modified Newtonian dynamics (MOND) paradigm \\citep{Mond3, mondnew} is known to reproduce galaxy scaling relations like the Tully-Fisher relation \\citep{tully}, the Faber-Jackson law \\citep{faber} and the fundamental plane \\citep{fundamental}) as well as the rotation curves of individual galaxies over five decades in mass \\citep{spiral1,mondref1,mondref2,mondref3,mondref4,mondref5,escape}. In particular, the recent kinematic analysis of tidal dwarf galaxies by \\cite{debris} is very hard to explain within the classical CDM framework while it is in accordance with MOND \\citep{tidal1,tidal2}. In addition, observations of a tight correlation between the mass profiles of baryonic matter and dark matter in relatively isolated (field) galaxies at all radii \\citep{insight2,insight} are most often interpreted as supporting MOND. Nevertheless, in rich clusters of galaxies, the MOND prescription is not enough to explain the observed discrepancy between visible and dynamical mass \\citep{neutrinos2,tevesfit,asymmetric}. At very large radii, the discrepancy is about a factor of $2$, meaning that there should be as much dark matter (mainly in the central parts) as observed baryons in MOND clusters. One solution is that neutrinos have a mass at the limit of detection, i.e. $\\sim2$ eV, which can solve the bulk of the problem of the missing mass in galaxy clusters, but other issues remain \\citep{group}. These $2$ eV neutrinos have also been invoked to fit the angular power spectrum of the cosmic microwave background (CMB) in relativistic MOND \\citep{tevesneutrinocosmo}, and are thus part of the only consistent MOND cosmology presented so far. In the following, we will refer to this model as the MOND hot dark matter ($\\mu$HDM) cosmology \\citep{tevesfit}\\footnotemark\\footnotetext[6]{Note, however, that one could also switch to sterile neutrinos with masses of a few eV \\citep[e.g.][]{sterile,maltoni1,maltoni2} and that massive (sterile) neutrinos are not indispensable within certain covariant formulations of modified gravity, e.g. the $V\\Lambda$ model, which can mimic the effects of neutrinos in clusters and cosmology as well as the behavior of a cosmological constant \\citep{vector}.}. On the other hand, strange features have recently been discovered in galaxy clusters, which are hard to explain, such as the ``dark matter core\" devoid of galaxies at the center of the ``cosmic train wreck\" cluster Abell 520 \\citep{abell520} and others \\citep{darkcluster,bullet}. Here, we consider the possibility that this kind of features could be due to the gravitational lensing effects generated by an intercluster filament in a universe based on tensor-vector-scalar gravity \\citep[TeVeS;][]{teves}, one possible relativistic extension of MOND \\citep[cf.][]{tv1,tv2,vector}. However, we are not performing a detailed lensing analysis of any particular cluster in the presence of filaments, but rather provide a proof of concept that the influence of filaments could be much less negligible in a MONDian universe than within the framework of general relativity (GR). Filaments are among the most prominent large-scale structures of the universe. From simulations in $\\Lambda$CDM cosmology, we know that almost every two neighboring clusters are connected by a straight filament with a length of approximately $20-30$ Mpc \\citep{LCDMfilament}. For instance, the dynamics of field galaxies, which are generally embedded in such filaments, as well as their weak lensing properties are persistently influenced by this kind of structures, generally encountering accelerations of about $0.01-0.1\\times 10^{-10}$ m s$^{-2}$. Filaments also cover a fair fraction of the sky, much larger than the covering factor of galaxy clusters. Thus, there is a good chance that filaments might be superimposed with other objects on a given line of sight, hence affecting the analysis of observational data like, for example, weak lensing shear measurements. Such recent studies prompted us to investigate the possibility that, in the context of MOND, end-on filamentary structures could be responsible for creating anomalous features in reconstructions of weak lensing convergence maps such as the peculiar ``dark matter core\" devoid of galaxies in Abell 520 \\citep{abell520}. Short straight filaments are structures which, at the best, are partially virialized in two directions perpendicular to their axis. According to \\cite{LCDMfilament}, a filament generally corresponds to an overdensity of about $10-30$, having a cigar-like shape. Furthermore, filamentary structures tend to have a low-density gradient along their axis and, in the perpendicular directions, they have a nearly uniform core which tapers to zero at larger radii, usually about $2-5$ times their core radius. Since filaments are typically much longer than their diameter, we shall approximately treat them as infinite uniform cylinders of radius $R_{f}=2.5$ $h^{-1}$ Mpc. Lacking a MOND/TeVeS structure formation $N$-body simulation (with or without substantially massive neutrinos), we shall adopt the naive assumption that filamentary structures have roughly the same properties in MOND and in CDM, which will be justified in \\S \\ref{app}. Deriving expressions for the TeVeS deflection angle and setting up a cosmological background, we conclude that the order of magnitude of the TeVeS lensing signal caused by filaments is compatible with that of the previously mentioned observed anomalous systems. In addition, we find that there is fundamental difference between GR and MOND/TeVeS for cylindrically symmetric lens geometries (see Fig. \\ref{fig1}); in contrast to GR, the framework of MOND/TeVeS allows us to have image distortion and amplification effects where the projected matter density is equal to zero. As for a more realistic approach, we also consider a model where the filament has a fluctuating density profile perpendicular to its axis. Compared to the uniform model, we find that the lensing signal in this case is smaller, but still of the same order, taking into account that the filamentary structures may be inclined to the line of sight by rather small angles ($\\theta\\lesssim 20^{\\circ}$). Finally, we demonstrate the impact of filaments onto the convergence map of other objects by considering superposition of such structures with a toy cluster along the line of sight. Again, our results show an additional contribution comparable to that of a single isolated filament. \\begin{figure} \\centering \\includegraphics[trim=-20 0 0 0,width=\\linewidth]{figures/Fig1.pdf} \\caption{Light deflection by an infinitely elongated cylinder of constant mass density; the unperturbed photon traveling along the $z$-direction passes the filament at the distance $y$ (impact parameter) from the filament's axis and is deflected by the angle $\\hat\\alpha$. The line density of the filament is assumed to be constant, $\\lambda = M/L = \\rho \\pi R_f^2$, where $\\rho$ is the volume density and $R_f$ is the cylinder's radius.} \\label{fig1} \\end{figure} ", "conclusions": "\\label{discussion} In this work, we have analyzed the gravitational lensing effect by filamentary structures in TeVeS, a relativistic formulation of the MOND paradigm. For this purpose, we have set up two different cosmological models in TeVeS: the so-called $\\mu$HDM cosmology including massive neutrinos on the order of $2$ eV which have already been proposed as a remedy for the discrepancies between dynamical and visible mass on cluster scales \\citep{neutrinos2} as well as for the CMB \\citep{tevesneutrinocosmo}, and a simple minimal-matter cosmology accounting for a universe which is made up of baryons alone. Encouraged by several HDM simulations and the fact that filamentary structures are generic, we have assumed that the properties of such structures, i.e. their shape and relative densities, are similar in CDM and MOND/TeVeS scenarios independent of the particularly used cosmological background. Modeling these filaments as infinite uniform mass cylinders, we have derived analytic expressions for their lensing properties in MOND/TeVeS and Newtonian/GR gravity. Regardless of the actual used cosmological background, we have shown that TeVeS filaments can account for quite a substantial contribution to the weak lensing convergence and shear field, $\\kappa \\sim \\gamma \\sim 0.01$, as well as to the amplification bias, $A^{-1}\\sim 1.02$, which is even true outside but close $(y\\sim 2R_{f})$ to the projected ``edges\" of the filament's matter density. Exploring a simple oscillating density model of a filament and its surrounding area, we have found that the lensing signal in this case is generally smaller, but can still be of the same order, taking into account that the filamentary structures may be inclined to the line of sight by rather small angles ($\\theta\\lesssim 20^{\\circ}$). In addition, we find that there is fundamental difference between GR and MOND/TeVeS considering idealized cylindrically symmetric lens geometries: wherever the projected matter density is zero, there will be no distortion as well as no amplification effects, i.e. image and source will look exactly the same. In the context of MOND/TeVeS, however, this changes as one can have such effects in these regions. Finally, we have demonstrated the impact of filaments onto the convergence map of other objects by considering superposition with a toy cluster along the line of sight. Again, our results have shown an additional contribution comparable to that of a single isolated filament and that the contribution pattern of filaments can be generally quite complex. Here we have considered the lensing signal generated by single filaments alone. Simulating the cosmic web in a standard $\\Lambda$CDM cosmology, \\cite{dolag} have found a shear signal $\\gamma\\sim 0.01-0.02$ along filamentary structures, which seems quite similar to what MOND/TeVeS can do. Note, however, that this signal is entirely dominated by the simulation's galaxy clusters, with the filament's signal being much smaller, approximately on the order of $10^{-4}-10^{-3}$. Although our analysis is mainly of theoretical interest, the above result points to an interesting possibility concerning recent measurements of weak lensing shear maps. For instance, the weak shear signal in the ``dark matter peak\" of Abell 520 \\citep{abell520} is roughly at a level of $0.02$, % which is comparable to what filaments could produce in MOND/TeVeS, but not in Newtonian gravity (also cf. \\cite{wedding}).% Therefore, we conclude that filamentary structures might actually be able to cause such anomalous lensing signals within the framework of MOND/TeVeS. In principle, the predicted difference in the weak lensing signal could also be used to test the validity of modified gravity. As several attempts to detect filaments by means of weak lensing methods have failed so far, e.g. the analysis of Abell $220$ and $223$ by \\cite{dietrich}, this might already be a first hint to possible problems within MOND/TeVeS gravity. On the other hand, shear signals around $\\gamma\\sim 0.01$ are still rather small to be certainly detected by today's weak lensing observations, and lacking $N$-body structure formation simulations in the framework of MOND/TeVeS, we cannot even be sure about how filaments form and how they look like in a MONDian universe compared to the CDM case. Clearly, more investigation is needed to gain a better understanding about the impact of filamentary structures." }, "0710/0710.3809_arXiv.txt": { "abstract": "The bipolar morphology of the planetary nebula (PN) K 3-35 observed in radio-continuum images was modelled with 3D hydrodynamic simulations with the adaptive grid code {\\sc yguaz\\'u-a}. We find that the observed morphology of this PN can be reproduced considering a precessing jet evolving in a dense AGB circumstellar medium, given by a mass loss rate $\\dot{M}_{csm}=5\\times 10^{-5}~\\mathrm{M_{\\odot} yr^{-1}}$ and a terminal velocity $v_\\mathrm{w}=10~\\mathrm{km s^{-1}}$. Synthetic thermal radio-continuum maps were generated from numerical results for several frequencies. Comparing the maps and the total fluxes obtained from the simulations with the observational results, we find that a model of precessing dense jets, where each jet injects material into the surrounding CSM at a rate $\\dot{M}_j=2.8\\times 10^{-4}~\\mathrm{M_{\\odot}\\ yr^{-1}}$ (equivalent to a density of $8 \\times 10^{4}~\\mathrm{cm}^{-3}$), a velocity of $1\\,500~\\mathrm{km~s^{-1}}$, a precession period of 100~yr, and a semi-aperture precession angle of 20\\degr\\ agrees well with the observations. ", "introduction": "\\label{sec:intro} The evolution between the end of the asymptotic giant branch (AGB) and the planetary nebula (PN) phases has for a long time been a poorly understood link in the late stages of intermediate-mass stars (1 -- 8 M$_\\odot$). It is in this transition phase where the fast stellar wind of the emerging PN interacts with the slow wind from its precursor AGB star (Kwok, Purton, \\& Fitzgerald 1978), shaping the final morphology of the PN. Other ingredients that can contribute to the final shape of a PN are the presence of interacting binary stars (e.g. Morris 1987; Balick \\& Frank 2002; Soker \\& Bisker 2006) or magnetic fields (e.g. Garc\\'\\i a-Segura \\& L\\'opez 2000; Blackman et al. 2001; Garc\\'\\i a-Segura, L\\'opez \\& Franco 2005). Thus, the study of objects in this transition phase can give us important clues about the physical mechanisms responsible for the different morphologies observed in PNe. K 3-35 is a very young PN with a characteristic S-shaped emission morphology that suggests the presence of precessing bipolar jets (Aaquist \\& Kwok 1989; Aaquist 1993; Miranda et al. 2000; 2001). The detection of OH and H$_2$O maser emission as well as the presence of CO and HCO$^+$ emission suggest that K~3-35 departed from the proto-PN phase only a few decades ago (Miranda et al. 2001; Tafoya et al. 2007). Miranda et al. (2001) estimate a dynamical age of $\\leq$ 15 years for the ionised inner core, which expands at $\\sim$25 km~s$^{-1}$. For the jets, assuming a modest jet velocity of $\\sim$100 km~s$^{-1}$, an age of $\\sim$800 years is obtained. Therefore, the jet formation in K~3-35 occurred during the proto-PN phase. In this paper we show that the radio morphology of K 3-35 can be explained by a precessing jet model. ", "conclusions": "Several 3D hydrodynamical simulations were carried out employing the adaptive grid code {\\sc yguaz\\'u-a}, in order to model both the morphology and the thermal radio emission of the planetary nebula K 3-35. After analysing our results, we found that the bipolar structure of this PN can be described as the result of the interaction of a dense jet (with an initial number density of $8\\sim 10^4$~cm$^{-3}$ or $\\dot{M}_j=2.8\\times 10^{-4} M_{\\odot} yr^{-1}$) moving into a dense environment, previously swept up by the AGB wind of the central star. The `S' morphology shown by K~3-35 in 8.3 GHz radio-continuum images \\citep{lfm98,lfm01,gomez03} can be reproduced if the modelled jet precesses with a period of 100~yrs on a cone with a half-opening angle of 20\\degr. For an integration time of 40~yrs, the simulated jet has a total length of $1.8\\times 10^{17}$~cm, which is equivalent to 2.4\\arcsec \\citep{lfm01,gomez03}considering an estimated distance of 5 kpc to K 3-35 and also the orientation of this object \\citep{uscanga07}. This time is almost 20 times smaller than the one given by \\citet{lfm01} for the jet, where a slower velocity was assumed. However, it is only $2.7$ times larger than the dynamical age of the inner core \\citep[also in][]{lfm01}, suggesting that the two events are more related than previously thought, and further supporting the idea that K 3-35 is a young object. Synthetic radio-continuum maps were generated from our numerical results. These maps show that the predicted morphologies and fluxes are in reasonable agreement with the observations. At an integration time of 40~yrs, the obtained spectral index is the one of optically thin emission. For an integration time of 50~yrs, the observed change of the spectral index with frequency (Aaquist 1993) is also reproduced by our simulation. A direct comparison between the observational and numerical results is given in Fig. \\ref{fcomp}, where we show the observed and synthetic (for an integration time of 40~yrs) 8.3 GHz radio maps. We must note that the values for the velocity and mass loss rate employed in the simulation for the jet and the CSM seem to be rather high. It is difficult for AGB and post-AGB starts to launch jets at velocities of the order of $1\\,500\\,\\mathrm{km~s}^{-1}$ (although Riera et al. 2003 have reported this kind of velocity for the outflows of the PPN 3-1475). Besides, the mass injection is also a bit extreme, in 40 years both jets have injected $0.02~\\mathrm{M}_{\\odot}$ into the surrounding CSM. This scenario might be explained in terms of a binary system, if the jet is produced by a companion accreting material (at a rate about ten times higher than $\\dot{M}_j$) from a massive star (the AGB progenitor that produced the density distribution of the circumstellar material). The primary in the last 40 years has lost $0.2~\\mathrm{M}_{\\odot}$. This means that it started with an envelope containing this mass, which appears to correspond to an AGB star rather than a post-AGB star (even though at the present time the star has already evolved to the planetary nebula stage). Furthermore, the CSM is very dense. With the parameters employed, in a radius of $10^{17}$~cm, a mass of 1 M$_{\\odot}$ is contained, implying a massive PN. There is observational evidence that favours a very dense surrounding CSM. However, the total mass derived from HCO$^+$ observations (Tafoya et al. 2007) is quite low ($\\sim 0.017~$M$_\\odot$), so that the molecular emission from this molecule would be confined to a small, possibly shock excited volume, in order to be consistent with our much higher mass CSM. These values imply that this scenario would be plausible if it is a short lived event. Clearly, a better determination of these parameters could be done with proper motion studies and better distance determinations. Notwithstanding the extreme values for the employed parameters, it is important to note that a simple model of a precessing dense jet moving into an also dense CSM, successfully reproduces the observed morphology of the PN K 3-35, obtaining a qualitatively and quantitatively good agreement between the model predictions and the observations, although this scenario would be a short lived event. \\begin{figure} \\includegraphics[width=\\columnwidth]{fig4.eps} \\caption{Comparison between observational and numerical results. The top panel shows the 8.3~GHz radio continuum image of K 3-35 \\citep{lfm01,gomez03}. The bottom panel shows the tilted synthetic radio emission map at the same frequency, obtained for an integration time of 40~yr. An inclination of 36\\degr between the precession axis and the plane of the sky was considered. The simulated map was smoothed with the observed beam of $0.2\\arcsec\\times 0.2\\arcsec$ and it is shown in the same scale that the observed one. The contours corresponds to the levels 0.18, 0.31, 0.56, 1.0, 1.78, 3.16, 5.62, and 10.00 mJy beam$^{-1}$. The arrows indicate the jet direction (projected on the plane of the sky) for integrations times of 0 and 40~yr. The last one is coincident with the P.A. measured by \\citet{lfm01}.} \\label{fcomp} \\end{figure}" }, "0710/0710.2722_arXiv.txt": { "abstract": "We present the results of the multi-frequency observations of radio outburst of the microquasar Cyg X-3 in February and March 2006 with the Nobeyama 45-m telescope, the Nobeyama Millimeter Array, and the Yamaguchi 32-m telescope. Since the prediction of a flare by RATAN-600, the source has been monitored from Jan 27 (UT) with these radio telescopes. At the eighteenth day after the quench of the activity, successive flares exceeding 1 Jy were observed successfully. The time scale of the variability in the active phase is presumably shorter in higher frequency bands. We also present the result of a follow-up VLBI observation at 8.4 GHz with the Japanese VLBI Network (JVN) 2.6 days after the first rise. The VLBI image exhibits a single core with a size of $<8$ mas (80 AU). The observed image was almost stable, although the core showed rapid variation in flux density. No jet structure was seen at a sensitivity of $T_b = 7.5\\times 10^5$ K\\@. ", "introduction": "Cyg X-3 is a famous X-ray binary including a black hole candidate (e.g., \\cite{schalinski1998}). This object is classified as a microquasar due to its bipolar relativistic jet accompanied by radio flares. Because it is located on the Galactic plane at a distance of about 10 kpc (e.g., \\cite{predehl2000}) and obscured by intervening interstellar matter, it has been observed mainly in radio and X-ray. Its giant radio flares have been observed once every several years since its initial discovery \\citep{gregory1972,braes1972}. The peak flux densities in the radio flares have often increased up to levels of 10 Jy or more at centimeter wave (e.g., \\cite{waltman1994}). The radio emission seems to be correlated with hard X-ray emission, and not with soft X-ray emission \\citep{mccollough1999}. Although the radio emission arises through synchrotron process of relativistic electrons in the jet \\citep{hjellming1988}, the millimeter behavior during the flares is not yet established. An observation at a shorter wavelength and with a higher time resolution is desirable to understand the mechanism of the flares. The quenched state of Cyg X-3, in which the radio emission is suppressed below 1 mJy, is a possible precursor of flares (e.g., \\cite{waltman1994}). In January 2006, this quenched state was detected in monitoring observations with the RATAN-600 radio telescope \\citep{trushkin2006}. The source has been monitored from MJD$=53762$ (Jan 27 2006 in UT) with the Nobeyama 45-m radio telescope (NRO45), the Nobeyama Millimeter Array (NMA), and the Yamaguchi 32-m radio telescope (YT32). We detected the initial state, or rising phase, of the radio flare of Cyg X-3 at MJD$= 53768$ (February 2 2006) and observed successive flares exceeding 1 Jy \\citep{tsuboi2006}, which turned out to be the beginning of an active phase lasting more than 40 days . In this paper, radio observations with NRO45, NMA, YT32, and the Japanese VLBI Network (JVN) are reported. The observation procedures are summarized in section 2. The light curves and the spectral evolution observed with NRO45, NMA and YT32 are shown in section 3, together with the result of JVN\\@. Detailed discussion based on these observations will be published as separate papers. ", "conclusions": "" }, "0710/0710.0727_arXiv.txt": { "abstract": "{ Gamma-ray burst are thought to be produced by highly relativistic outflows. Although upper and lower limits for the outflow initial Lorentz factor $\\Gamma_0$ are available, observational efforts to derive a direct determination of $\\Gamma_0$ have so far failed or provided ambiguous results. As a matter of fact, the shape of the early-time afterglow light curve is strongly sensitive on $\\Gamma_0$ which determines the time of the afterglow peak, i.e. when the outflow and the shocked circumburst material share a comparable amount of energy. We now comment early-time observations of the near-infrared afterglows of GRB\\,060418 and GRB\\,060607A performed by the REM robotic telescope. For both events, the afterglow peak was singled out and allowed us to determine the initial fireball Lorentz, $\\Gamma_0\\sim400$. ", "introduction": "The early stages of gamma-ray burst (GRB) afterglow light curves display a rich variety of phenomena at all wavelengths and contain significant information which may allow determining the physical properties of the emitting fireball. The launch of the \\textit{Swift} satellite \\citep{Neil04}, combined with the development of fast-slewing ground-based telescopes, has hugely improved the sampling of early GRB afterglow light curves. Since many processes work in the early afterglow, it is often difficult to model them well enough to be able to determine the fireball characteristics. The simplest case is a light curve shaped by the forward shock only. This case is particularly interesting because, while the late-time light curve is independent of the initial conditions (the so-called self-similar solution), the time at which the afterglow peaks depends on the original fireball Lorentz factor $\\Gamma$, thus allowing a direct measurement of this fundamental parameter \\citep{SP99}. The short variability timescales, coupled with the nonthermal GRB spectra, indeed imply that the sources emitting GRBs have a highly relativistic motion \\citep{Ruderman75,Fenimore93,Piran00,Lith01}, to avoid suppression of the high-energy photons due to pair production. This argument, however, can only set a lower limit to the fireball Lorentz factor. Late-time measurements (weeks to months after the GRB) have shown $\\Gamma \\sim \\mbox{a few}$ \\citep{Frail97,Taylor05}, but a direct measure of the initial value (when $\\Gamma$ is expected to be $\\sim 100$ or more) is still lacking. We present here the NIR early light curves of the GRB\\,060418 and GRB\\,060607A afterglows observed with the REM robotic telescope\\footnote{\\url{http://www.rem.inaf.it}} \\citep{Zerb01,Chinc03} located in La Silla (Chile). These light curves show the onset of the afterglow and its decay at NIR wavelengths as simply predicted by the fireball forward shock model, without the presence of flares or other peculiar features. A detailed discussion of these data has also been reported by \\citet{Mol07} and \\citet{Mal07}. \\begin{figure} \\centering \\includegraphics[width=11cm]{CovinoS_2007_01_fig01.ps} \\\\ \\centering\\includegraphics[width=11cm]{CovinoS_2007_01_fig02.ps} \\caption{NIR and X-ray light curves of GRB\\,060418 (upper panel) and GRB\\,060607A (lower panel). The dotted lines show the models of the NIR data using the smoothly broken power law (see Sect. \\ref{modelling}). For GRB\\,060418 the dashed line shows the best-fit to the X-ray data.\\label{fig:lcs}} \\end{figure} ", "conclusions": "The REM discovery of the afterglow onset has demonstrated once again the richness and variety of physical processes occurring in the early afterglow stages. The very fast response observations presented here provide crucial information on the GRB fireball parameters, most importantly its initial Lorentz factor. This is the first time that $\\Gamma(t_{\\rm peak})$ is directly measured from the observations of a GRB. The measured $\\Gamma_0$ value is well within the range ($50 \\la \\Gamma_{0} \\la 1000$) envisaged by the standard fireball model \\citep{Piran00,Guetta01,Alicia02,Meszaros}. It is also in agreement with existing measured lower limits \\citep{Lith01,zhang06}. Using $\\Gamma_0=400$ we can also derive other fundamental quantities characterising the fireball of the two bursts. In particular, the isotropic-equivalent baryonic load of the fireball is $M_{\\rm fb} = E/(\\Gamma_0 c^2) \\approx 7\\times 10^{-4}\\,M_\\odot$, and the deceleration radius is $R_{\\rm dec} \\approx 2 c t_{\\rm peak} [\\Gamma(t_{\\rm peak})]^2/(1+z) \\approx 10^{17}$~cm. This is much larger than the scale of $\\sim 10^{15}$~cm where the internal shocks are believed to power the prompt emission \\citep{MesRees97,Rees94}, thus providing further evidence for a different origin of the prompt and afterglow stages of the GRB." }, "0710/0710.0382_arXiv.txt": { "abstract": "% We present results from a 150 ksec {\\it Suzaku} observation of the Seyfert 1.5 NGC 3516 in October 2005. The source was in a relatively highly absorbed state. Our best-fit model is consistent with the presence of a low-ionization absorber which has a column density near 5 $\\times$ 10$^{22}$ cm$^{-2}$ and covers most of the X-ray continuum source (covering fraction 96--100$\\%$). A high-ionization absorbing component, which yields a narrow absorption feature consistent with Fe K {\\sc XXVI}, is confirmed. A relativistically broadened Fe K$\\alpha$ line is required in all fits, even after the complex absorption is taken into account; an additional partial-covering component is an inadequate substitute for the continuum curvature associated with the broad Fe line. A narrow Fe K$\\alpha$ emission line has a velocity width consistent with the Broad Line Region. The low-ionization absorber may be responsible for producing the narrow Fe K$\\alpha$ line, though a contribution from additional material out of the line of sight is possible. We include in our model soft band emission lines from He- and H-like ions of N, O, Ne and Mg, consistent with photo-ionization, though a small contribution from collisionally-ionized emission is possible. ", "introduction": "The 6.4 keV Fe K$\\alpha$ emission line has long been known to be an important diagnostic of the material accreting onto supermassive black holes. The Compton reflection hump, frequently seen in Seyfert spectra above $\\sim$7~keV and peaking near 20--30 keV (Pounds et al.\\ 1990), indicates that Seyferts' Fe K lines may have an origin in optically thick material, such as the accretion disk. Observations with {\\it ASCA} indicated that many Fe K$\\alpha$ lines were broad (FWHM velocities up to $\\sim$0.3$c$) and asymmetrically skewed towards lower energies, implying an origin in the inner accretion disk; the line profile is sculpted by gravitational redshifting and relativistic Doppler effects (e.g., Tanaka et al.\\ 1995, Fabian et al.\\ 2002). However, {\\it XMM-Newton} and {\\it Chandra} observations have been revealing a more complex picture. A narrow Fe K component (FWHM velocities $\\sim$5000 km s$^{-1}$ or less) appears to be quite common; these lines' widths suggest emission from distant material, such as the outer accretion disk, the optical/UV Broad Line Region (BLR) or the molecular torus. Spectral observations in which the broad and narrow components are deconvolved are thus a prerequisite for using the Fe K line as a tracer of the geometry of the emitting gas. At the same time, there is strong evidence from X-ray and UV grating observations for the presence of ionized material in the inner regions of a large fraction of AGN (e.g., Blustin et al.\\ 2005; McKernan et al.\\ 2007). High-resolution spectroscopy shows the gas is usually outflowing from the nucleus; typical velocities are $\\sim$ a few hundred km s$^{-1}$. Absorption due to a broad range of ionic species is commonly seen; and for many sources, there is evidence for several different photo-ionized absorbing components, as opposed to a single absorber, along the line of sight. In the Fe K bandpass, some Seyferts show evidence for absorption by H- or He-like Fe, indicating a zone of highly-ionized absorbing material (e.g., NGC 3783, Reeves et al.\\ 2004). Cold absorbing gas, with line of sight columns in excess of the Galactic value, are routinely observed in Seyfert 2 AGN in accordance with unification schemes (Urry, Padovani 1995), and have also been reported in some Seyfert 1 AGN. Importantly, variations in column density on timescales from hours to years have been observed in both Seyfert 1 AGN (e.g., in I Zw 1, Gallo et al.\\ 2007; see also Lamer et al.\\ 2003, Puccetti et al.\\ 2004) and Seyfert 2 AGN (Risaliti et al.\\ 2002; Risaliti et al.\\ 2005), suggesting that the absorbing circumnuclear material is not homogeneous in either Seyfert type, has a high tranverse velocity and occurs over a range of length scales. NGC 3516 is a well-studied, nearby (z = 0.008836; Keel 1996) Seyfert 1.5 AGN that can display strong 2--10 keV flux variability on timescales of hours to years (e.g., Markowitz, Edelson 2004). Previous X-ray spectroscopic observations of NGC 3516 e.g., Nandra et al.\\ (1997) using {\\it ASCA}, have indicated the presence of a broad Fe K line, and this source is known to also contain complex and ionized absorption. Numerous UV absorption lines, including N {\\sc V}, C {\\sc IV} and Si {\\sc IV}, were observed with the {\\it International Ultraviolet Explorer} (Ulrich, Boisson 1983); absorption line strengths vary on timescales as short as weeks as the absorber responds to variations in the ionizing flux (e.g., Voit et al.\\ 1987). {\\it Hubble Space Telescope} observations have revealed that this component of absorbing gas (henceforth called the ``UV absorber'') may consist of several distinct kinematic components (Crenshaw et al.\\ 1998). X-ray spectra of NGC 3516 can exhibit evidence for large columns ($\\gtrsim$10$^{22}$ cm$^{-2}$) of absorbing gas (e.g., Kolman et al.\\ 1993), and the X-ray absorbers can display variability on timescales of years (e.g., Mathur et al.\\ 1997). Using {\\it Chandra} gratings data from observations in April 2001 and November 2001, Turner et al.\\ (2005) observed K-shell absorption lines due to H- like Mg, Si and S, and He-like Si, evidence for a highly-ionized absorber, likely with column density $\\gtrsim$10$^{22}$ cm$^{-2}$, outflowing at $\\sim$1100 km s$^{-1}$. Simultaneous with these {\\it Chandra} observations in 2001 were two {\\it XMM-Newton} observations. Turner et al.\\ (2005) modeled the continuum curvature of the two {\\it XMM-Newton} EPIC spectra by including a partial covering, mildly-ionized absorber; the column density was $\\sim$2.5 $\\times$ 10$^{23}$ cm$^{-2}$, with a covering fraction of $\\sim$50$\\%$. However, the formal requirement for the broad Fe line was reduced, leading to uncertainty as to whether the broad Fe line really existed in NGC 3516. Spectral fitting using an instrument with a wide bandpass is thus necessary to remove such model degeneracies. In this paper, we report on an observation of NGC 3516 made with the {\\it Suzaku} observatory in October 2005. The combination of the X-ray Imaging Spectrometer (XIS) CCD and the Hard X-ray Detector (HXD) instruments have yielded a broadband spectrum covering 0.3 to 76 keV, allowing us to deconvolve the various broadband emitting and absorbing components. Furthermore, the exceptional response of the XIS CCD and high signal-to-noise ratio of this observation have allowed us to study narrow emission lines in great detail. $\\S$2 gives a brief overview of the {\\it Suzaku} observatory, and describes the observation and data reduction. $\\S$3 describes fits to the time-averaged spectrum. Variability analysis is briefly discussed in $\\S$4. Flux-resolved spectral fits are discussed in $\\S$5. In $\\S$6, we describe a search for narrow red- or blue-shifted lines in the Fe K bandpass. The results are discussed in $\\S$7, and a brief summary is given in $\\S$8. ", "conclusions": "During the late 1990's, NGC 3516 typically displayed a 2--10 keV flux of $\\sim$4--6 $\\times$ 10$^{-11}$ erg cm$^{-2}$ s$^{-1}$ (e.g., Markowitz, Edelson 2004). During the 2001 {\\it XMM-Newton}/{\\it Chandra} observations, however, the observed 2--10 keV flux was much lower: 1.6--2.3 $\\times$ 10$^{-11}$ erg cm$^{-2}$ s$^{-1}$ (Turner et al.\\ 2005). Table 4 lists the inferred absorption-corrected 2--10 keV nuclear fluxes from the {\\it XMM-Newton} observations, as well as during the 2005 {\\it Suzaku} observation. In addition, Figure 10 shows the unfolded observed spectra for the {\\it Suzaku} XIS and the 2001 {\\it XMM-Newton} EPIC-pn data (see Turner et al.\\ 2005 for details regarding the {\\it XMM-Newton} data). The {\\it Suzaku} observation apparently caught the source in a similar low level of nuclear flux as the 2001 observations. The observed 0.5--2.0 keV flux during the {\\it Suzaku} observation, however, was $\\sim$2--3 times lower than during the {\\it XMM-Newton} observations, indicating that the source was still heavily obscured, and confirming that the complex absorption in this source cannot be ignored when fitting the broadband spectrum and modeling diskline components. \\subsection{Power-law Components} The primary power-law observed in the hard X-rays is likely emission from a hot corona very close to the supermassive black hole, as seen in all Seyferts. The nature of the soft power-law component, however, is not as clear. It could represent nuclear emission scattered off optically-thin material, e.g., in the optical/UV Narrow Line Region (NLR). In the baseline model, the normalization of the soft power-law relative to that of the primary power-law was 4.2$\\pm$0.4$\\%$. Assuming a covering fraction of unity, this ratio is equal to the optical depth of the scattering material, indicating a column density of roughly 5 $\\times$ 10$^{22}$ cm$^{-2}$, consistent with this notion, though the column density is somewhat too high to likely be associated with the NLR. It is interesting to note that this column density is similar to that obtained for the high-ionization absorber, suggesting the possibility that this absorbing component could be associated with a zone of scattering. Using {\\it Chandra}-ACIS, George et al.\\ (2002) found the extended circumnuclear gas to have a 0.4--2.0 keV flux of roughly 10$^{-14}$ erg cm$^{-2}$ s$^{-1}$. However, that emission was studied over an annular extraction region 3$\\arcsec$ to 10$\\arcsec$ (0.6--1.8 kpc), and so that flux value is likely a lower limit to the 0.4--2.0 keV flux that Suzaku would observe (given the XRTs' 2$\\arcmin$ HPD). In our baseline model, we found an unabsorbed 0.4--2.0 keV flux of 1 $\\times$ 10$^{-12}$ erg cm$^{-2}$ s$^{-1}$, consistent with the notion that the soft power-law is scattered emission. In this case, the decrease in observed 0.5--2.0 keV flux from 2001 to 2005 could potentially be explained by the scattered emission responding to a recent decrease in nuclear continuum flux. However, this scenario would require the bulk of the scattered emission to lie within at most a few light years of the black hole, and the nuclear flux would have had to decrease between 2001 and 2005 (when the source was not observed by any major X-ray mission in the 2--10 keV band) then return to 2001 levels by the {\\it Suzaku} observation. Alternatively, the soft power-law could be unobscured, ``leaked'' nuclear emission as part of a partial covering scenario. In this case, the primary absorber would obscure 96$\\%$ of the sky as seen from the nuclear continuum source. The lack of significant variability in the 0.3--1.0 keV band could argue for the soft power-law to originate in scattered emission, since we might expect to observed variability of the same amplitude as the 2--10 keV band only if the soft power-law were leaked nuclear emission. However, this is far from certain, as the 0.3--1.0 keV band had a low count rate and the presence of the soft emission lines in the XIS spectrum could contribute to dilution of intrinsic variability in the soft power-law. A broadband observation spanning a larger observed flux range is needed to clarify this issue. The soft power-law could of course represent a blend of scattered emission plus leaked nuclear emission. We therefore conclude that the primary absorber has a covering fraction between 96--100$\\%$. \\subsection{Complex Absorption} We detect two zones of absorption: in addition to the primary absorber, which has a covering fraction of 96--100$\\%$, there is the high-ionization absorber, which is assumed here to have a covering fraction of unity. The high-ionization absorber is potentially the same as that reported by Turner et al.\\ (2005); we find a column density $N_{\\rm H}$ of 4.0$^{+4.6}_{-3.1}$ $\\times$ 10$^{22}$ cm$^{-2}$, consistent with the column density of 2 $\\times$ 10$^{22}$ cm$^{-2}$ used by Turner et al.\\ (2005), although we use a slightly higher ionization parameter (see Turner et al.\\ 2007). Previous studies of NGC 3516, such as Netzer et al.\\ (2002), have discussed in detail the UV absorber, responsible for H Ly$\\alpha$, C {\\sc IV} and N {\\sc V} absorption features in {\\it Hubble Space Telescope} spectra (Kraemer et al.\\ 2002). In the X-ray band, discrete features associated with Mg {\\sc VII--IX} and Si {\\sc VII--IX} are expected from this component, but with the CCD resolution and with calibration-related artifacts near 1.7--1.8 keV in the XIS, such features are not detected by {\\it Suzaku}. {\\it Suzaku} has found the primary absorber of the hard X-ray continuum to be lowly-ionized (log($\\xi$) = 0.3$\\pm$0.1 erg cm s$^{-1}$), with a column density $N_{\\rm H}$ of 5.5$\\pm$0.2 $\\times$ 10$^{22}$ cm$^{-2}$. It is possible that it is the same absorber that Turner et al.\\ (2005) designated as the ``heavy'' partial-covering absorber, though we use a somewhat lower ionization parameter (see Turner et al.\\ 2007). A new observation of NGC 3516 with {\\it XMM-Newton} in 2006 October showed that the source had returned to similar $>$6 keV brightness and similar obscuration levels ($N_{\\rm H}$ $\\sim$ 2 $\\times$ 10$^{23}$ cm$^{-2}$; covering fraction $\\sim$ 45$\\%$) as during the 2001 observations (Turner et al.\\ 2007). Long-term changes in the covering fraction of the heavy absorber could explain the bulk of the spectral variability changes between the 2001, 2005 and 2006 observations. In this case, the column density has decreased by a factor of 4.5, while the covering fraction has increased from $\\sim$40--60$\\%$ to 96--100$\\%$, from 2001 to 2005, and subsequently returned to approximately the same levels as in 2001 within the next 12 months. However, because we have not actually observed the entire eclipse associated with a specific, discrete blob of absorbing gas traversing the line of sight, it is not clear whether the covering fractions derived are associated with single, large blobs partially blocking the line of sight to the X-ray continuum source, or if the absorber consists of numerous, discrete blobs or has a filamentary or patchy structure. On the other hand, given the gaps between the 2001 and 2006 {\\it XMM-Newton} and 2005 {\\it Suzaku} observations, it is certainly plausible that these observations could have caught independent, discrete blobs or filaments with differing column densities and differing physical sizes and/or radial distances lying on the line of sight. To estimate the distance $r$ between the central black hole and the absorbing gas, we can use a definition of the ionization parameter $\\xi$ = $L_{1-1000 {\\rm Ryd}}$/($n$$r^2$), where $n$ is the number density. $L_{1-1000 {\\rm Ryd}}$ is the 1--1000 Ryd illuminating continuum luminosity, and the value of the ionization parameter is taken to be the value of 2 erg cm s$^{-1}$ measured above. We estimate the maximum possible distance to the material by assuming that the thickness $\\Delta$$r$ must be less than the distance $r$. The column density $N_{\\rm H}$ = $n$$\\Delta$$r$, yielding the upper limit $r$ $<$ $L_{1-1000 {\\rm Ryd}}$/($N_{\\rm H}$$\\xi$). We estimate the 1-1000 Ryd flux from the baseline model to be 9.9 $\\times$ 10$^{-11}$ erg cm$^{-2}$ s$^{-1}$, which corresponds to $L_{1-1000 {\\rm Ryd}} = 1.7 \\times 10^{43}$ erg s$^{-1}$ (assuming $H_{\\rm o}$ = 70 km s$^{-1}$ Mpc$^{-1}$ and $\\Lambda_{\\rm o}$ = 0.73). $r$ is thus $<$2 $\\times$ 10$^{20}$ cm (180 light-years), a very loose upper limit encompassing both distances associated with the BLR ($\\sim$10 light-days; Peterson et al.\\ 2004) and a possible cold molecular torus at a 1 pc radius. Variability in the absorber properties between 2001--2005 and 2005--2006 thus imply a radial distance of at most a few light years in the case of clouds traversing the line of sight to the nucleus. In addition, in the case of a partial covering scenario, it is plausible that the absorber's size could be of the same order as that of the X-ray continuum source. The absorbing material could thus be e.g., associated with the base of an outflow or from dense clouds associated with magnetohydrodynamic disk turbulence (e.g., Emmering et al.\\ 1992). NGC 3516's transition from an unobscured source to a moderately-obscured source in a 4 year span presents a challenge to standard Seyfert 1/2 unification schemes. If the obscuration in NGC 3516 is associated with an equatorial molecular torus usually invoked in Seyfert 1/2 unification schemes, then it is possible that during the {\\it Suzaku} observation, the inner edge of the torus could have intersected the line of sight, but given NGC 3516's classification as a Seyfert 1, this could only occur if the torus opening angle were extremely small and the torus were not azimuthally symmetric. Alternatively, variations in column density and/or covering fraction could be due to fine structure in large-distance (tens of pc), non-equatorial filaments that traverse the line of sight (e.g., Malkan et al.\\ 1998). \\subsection{Fe K Emission Components and Compton Reflection} We have deconvolved the broadband emitting components, and determined that 1) the existence of the broad Fe line is robust in that it was required in all models for an adequate fit, and 2) a partial covering component could not mimic the curvature associated with a relativistic broad line. We note, for instance, that if we remove the diskline components from the baseline model and refit, not only is the fit worse ($\\chi^2$ increases by over 170), but the value of $R$ becomes $\\sim$ 3.2. This value is incompatible with the observed $EW$ of the narrow line unless the Fe abundance is extremely sub-solar ($\\lesssim$0.3; see below). The best-fit disk inclination was typically $\\lesssim$25$\\arcdeg$. The inner radius was typically $\\lesssim$5$R_{\\rm g}$. The line energy was seen to be consistent with neutral to mildly-ionized Fe (up to Fe $\\sim$ {\\sc XX}; Kallman et al.\\ 2004). The equivalent width with respect the primary continuum was 287$^{+49}_{-24}$ eV, consistent with the value of 431$^{+193}_{-172}$ eV obtained by Turner et al.\\ (2005) for the April 2001 {\\it XMM-Newton} observation, where the spectrum could be fit with a diskline component in addition to the complex absorbing components. The line energy of the narrow Fe K$\\alpha$ line was also consistent with emission from neutral Fe. The intensity of the narrow line during the {\\it Suzaku} observation is roughly 40$\\%$ higher than that measured during the 2001 {\\it XMM-Newton} and {\\it Chandra} observations (Turner et al.\\ 2002), possibly indicating that a substantial fraction of the Fe K line photons originate in a region $\\lesssim$5 lt.-yr.\\ in size. We measured a FWHM velocity line width for the narrow Fe K$\\alpha$ line of $<$ 4900 km s$^{-1}$ (99$\\%$ confidence level for two interesting parameters). This velocity does not rule out an origin in the BLR; Peterson et al.\\ (2004) reported FWHM velocities for the H$\\alpha$ and H$\\beta$ lines of 4770$\\pm$893 and 3353$\\pm$310 km s$^{-1}$, respectively. However, we also cannot exclude a contribution from an origin in the putative molecular torus; there could potentially be a very narrow line component with FWHM velocity $\\sim$ a few hundred km s$^{-1}$, but the XIS would be unable to separate it from the relatively broader line component. It is possible that the same material that absorbs the hard X-rays along the line of sight is responsible for producing the narrow Fe line. The material producing the Fe line cannot have a column substantially less than 10$^{\\sim 22}$ cm$^{-2}$ or else there would be insufficient optical depth to produce a prominent Fe K line. The primary absorber, with its column density of 5.5 $\\times$ 10$^{22}$ cm$^{-2}$ and low ionization state, is thus a plausible candidate for the narrow line origin. As an estimate of the Fe K$\\alpha$ equivalent width expected in this case, we can assume an origin in optically-thin gas which completely surrounds a single X-ray continuum source and is uniform in column density, and use the following equation: \\begin{equation} EW_{\\rm calc} = f_{\\rm c} \\omega f_{\\rm K\\alpha} A \\frac{\\int^{\\infty}_{E_{\\rm K edge}}P(E) \\sigma_{\\rm ph}(E) N_{\\rm H} dE}{P(E_{\\rm line})} \\end{equation} Emission is assumed to be isotropic. Here, $f_{\\rm c}$ is the covering fraction, initially assumed to be 1.0. $\\omega$ is the fluorescent yield, 0.34 (Kallman et al.\\ 2004). $f_{\\rm K\\alpha}$ is the fraction of photons that go into the K$\\alpha$ line as opposed to the K$\\beta$ line; this is 0.89 for Fe {\\sc I}. $A$ is the number abundance relative to hydrogen. We assumed solar abundances, using Lodders (2003). $P(E)$ is the spectrum of the illuminating continuum at energy $E$; $E_{\\rm line}$ is the K$\\alpha$ emission line energy. $\\sigma_{\\rm ph}(E)$ is the photo-ionization cross section assuming absorption by K-shell electrons only (Veigele 1973\\footnote{http://www.pa.uky.edu/$\\sim$verner/photo.html}). For $N_{\\rm H}$ = 5.5 $\\times$ 10$^{22}$ cm$^{-2}$, $EW_{\\rm calc}$ = 29 eV, substantially lower than the observed $EW$ of 123$\\pm$7 eV. We conclude that it is possible for the primary absorber to contribute to the observed line $EW$, but there is also likely a contribution from some other (non-continuum absorbing) material lying out of the line of sight, likely with column densities 10$^{\\sim 23}$ cm$^{-2}$ (e.g., Matt et al.\\ 2002). For instance, if the putative cold molecular torus does not intersect the line of sight, it could contribute to the observed $EW$. The 13$\\%$ upper limit to ratio of the Compton shoulder/ narrow Fe K$\\alpha$ core intensity was a 90$\\%$ confidence limit only, and does not exclude at high confidence the possibility of Compton-thick material out of the line of sight. An additional possibility is that the material emitting the bulk of the line photons could be responding to a continuum flux that was higher in the past. For instance, if the putative molecular torus is located $\\sim$ a pc or so from the black hole, the torus will yield a line $EW$ corresponding to the continuum flux averaged over the past few years. This situation is plausible for NGC 3516, as the 2--10 keV flux of NGC 3516 during $\\sim$1998--2001 (Markowitz, Edelson 2004) was a factor of $\\sim$1.5--2 times brighter than during 2005. We now turn our attention to properties of the Compton reflection continuum. {\\it Suzaku} has observed other Seyferts to display reduced levels of variability in the PIN band compared to the 2--10 keV band, e.g., in MCG--6-30-15 (Miniutti et al.\\ 2007). This behavior is thought to be caused by the presence of the relatively non-varying Compton reflection hump, which dilutes the observed $>$10 keV variability of the power-law component. Gravitational light-bending in the region of strong gravity has been invoked to explain the relative constancy of the reflection spectrum (Compton hump and Fe K diskline component) despite large variations in the coronal power-law flux in MCG--6-30-15, for instance (Miniutti et al.\\ 2007). In the case of NGC 3516, the observed fractional variability amplitudes for the 2--10 and 12--76 keV bands were $F_{\\rm var,2-10}$ = 9.2$\\pm$0.3$\\%$ and $F_{\\rm var,12-76}$ $<$ 4.4$\\%$, respectively. These measurements allow us to rule out the possibility that the Compton hump varies in concert with the power-law, since the variability amplitudes would be consistent in that case. The primary power-law and Compton hump contribute 44$\\%$ and 56$\\%$, respectively, of the total 12--76 keV flux. In the case of a constant Compton hump and variable power-law, $F_{\\rm var,12-76}$ would then be equal to $F_{\\rm var,2-10}$ / 2.25, or roughly 4.1$\\%$. This is roughly consistent with the observed upper limit on $F_{\\rm var,12-76}$, suggesting that the reflection component varies less strongly than the primary power-law over the course of the observation. To verify this, however, we would need to observe NGC 3516 over a larger % X-ray flux range than in the current {\\it Suzaku} observation to potentially observe any significant variability in the PIN band. Finally, we discuss the origin of the material that gives rise to the observed Compton reflection hump. The primary and high-ionization absorbers lack the necessary column density, and are excluded. We next consider an origin in the same material that yields either the broad or narrow Fe lines. George, Fabian (1991) calculated that $R$ = 1 corresponds to an observed Fe K$\\alpha$ line $EW$ (relative to a primary continuum with a photon index of 1.9) of 140 eV for neutral Fe, assuming an inclination angle of 25$\\arcdeg$. However, George, Fabian (1991) used the elemental abundances of Morrison, McCammon (1983), where the Fe number abundance relative to hydrogen was $A_{\\rm Fe}$ = 3.3 $\\times$ 10$^{-5}$. More recent papers have slightly lower values of $A_{\\rm Fe}$, 2.7--3.0 $\\times$ 10$^{-5}$ (Lodders 2003; Wilms et al.\\ 2000). The expected equivalent width corresponding to $R$ = 1 is thus 115--125 eV. In our baseline model, we found a best-fit value of $R$ = 1.7, which corresponds to an expected line $EW$ (relative to the primary continuum) of 200--215 eV, a value in between the observed $EW$s of the broad line (287 eV in the baseline model) and the narrow line (123 eV). It is thus not clear from this measurement alone whether the total Compton reflection continuum is associated with the broad line (disk), narrow line (a distant origin), or both. That is, while is it a possibility that at least some portion of the Compton reflection component is associated with the broad Fe K component, we cannot exclude the possibility that the narrow line contributes as well and that there is reflection off cold, distant material. For example, in $\\S$3.3, we demonstrated that a model wherein there existed both blurred reflection from an ionized disk plus reflection from cold, distant material, such as the molecular torus, provided a good fit to the data. In addition, we demonstrated in this section that the observed $EW$ of the narrow Fe K line means we cannot rule out a contribution to the narrow Fe line, and to the reflection continuum as well, from Compton-thick material out of the line of sight." }, "0710/0710.2343.txt": { "abstract": "This paper analyses the radio properties of a subsample of optically obscured ($R\\ge 25.5$) galaxies observed at 24$\\mu$m by the {\\it Spitzer Space Telescope} within the First Look Survey. 96 $F_{24\\mu\\rm m}\\ge 0.35$~mJy objects out of 510 are found to have a radio counterpart at 1.4~GHz, 610~MHz or at both frequencies respectively down to $\\sim 40\\mu$Jy and $\\sim 200\\mu$Jy. IRAC photometry sets the majority of them in the redshift interval $z\\simeq [1-3]$ and allows for a broad distinction between AGN-dominated galaxies ($\\sim 47$\\% of the radio-identified sample) and systems powered by intense star-formation ($\\sim 13$\\%), the remaining objects being impossible to classify. The percentage of radio identifications is a strong function of 24$\\mu$m flux: almost all sources brighter than $F_{24\\mu\\rm m}\\sim 2$~mJy are endowed with a radio flux at both 1.4~GHz and 610~MHz, while this fraction drastically decreases by lowering the 24$\\mu$m flux level. The radio number counts at both radio frequencies suggest that the physical process(es) responsible for radio activity in these objects have a common origin regardless of whether the source shows mid-IR emission compatible with being an obscured AGN or a star-forming galaxy. We also find that both candidate AGN and star-forming systems follow (although with a large scatter) the relationship between 1.4~GHz and 24$\\mu$m fluxes reported by Appleton et al. (2004) which identifies sources undergoing intense star formation activity. However, a more scattered relation is observed between 24$\\mu$m and 610~MHz fluxes. On the other hand, the inferred radio spectral indices $\\alpha$ indicate that a large fraction of objects in our sample ($\\sim 60$\\% of all galaxies with estimated $\\alpha$) may belong to the population of Ultra Steep Spectrum (USS) Sources, typically 'frustrated' radio-loud AGN. We interpret our findings as a strong indication for concurrent AGN and star-forming activity, whereby the 1.4~GHz flux is of thermal origin, while that at 610~GHz mainly stems from the nuclear source. ", "introduction": "The advent of the {\\it Spitzer Space Telescope} has marked a fundamental milestone in our understanding of the assembly history of massive spheroidal galaxies, one of the major issues for galaxy formation models. The unprecedented sensitivity of the Multiband Imaging Photometer for {\\it Spitzer} (MIPS) at 24$\\mu$m has in fact for the first time allowed the detection at high redshifts of a population of Luminous and UltraLuminous Infrared Galaxies (LIRGs; ULIRGs) with huge infrared luminosities ($L_{\\rm IR}>10^{11}L_{\\odot}$). Such sources are underluminous at rest frame optical and UV wavelengths because they are reprocessing and radiating much of their energy in the IR (e.g. Sanders \\& Mirabel 1996). As a consequence, LIRGs and ULIRGs at high redshifts had been missed so far either due to their extreme optical faintness or because previous infrared missions such as the InfraRed Astronomical Satellite (IRAS) or the Infrared Space Observatory (ISO) did not have enough sensitivity to push the observations beyond $z \\sim 1$. Recent studies have shown that these objects, while relatively rare in the local universe (e.g. Sanders \\& Mirabel 1996), become an increasingly significant population at higher redshifts (e.g. Le Floc'h et al. 2004; 2005; Lonsdale et al. 2004; Caputi et al. 2006; 2007) and likely dominate the luminosity density at $z>1$ (see e.g. Dole et al. 2006). Their space density, found to range between $10^{-4}$ and a few $10^{-5}$~Mpc$^{-3}$ according to the selection criteria adopted by different studies (see e.g. Caputi et al 2007; Daddi et al. 2007a, Magliocchetti et al. 2007a), is a factor of 10 to 100 higher than that of optically selected quasars in the same redshift range (e.g. Porciani, Magliocchetti \\& Norberg 2004). Furthermore, clustering studies (e.g. Magliocchetti et al. 2007; 2007a; Farrah et al. 2006) prove that, at variance with their local counterparts, LIRGs and ULIRGs at $z\\sim 2$ are associated with extremely massive ($M\\simgt 10^{13} M_{\\odot}$, where $M$ here refers to the dark matter) structures, only second to those which locally host very rich clusters of galaxies. Given their properties, it then appears clear that these sources represent a fundamental phase in the build up of massive galactic bulges, and in the growth of their supermassive black holes. Emission line diagnostics for bright ($F_{24\\mu \\rm m}\\simgt 1$~mJy) mid-IR samples of LIRGs and ULIRGs at $z\\sim 2$ in the near and mid-IR spectral regimes (see e.g. Yan et al. 2005; 2007; Weedman et al. 2006; 2006a; Brand et al. 2007; Martinez-Sansigre et al. 2006a) have shown these sources to be a mixture of obscured type1-type2 AGN and systems undergoing intense star formation activity. These findings are confirmed by photometric follow up mainly undertaken in the mid-IR and X-ray (both soft and hard) bands which also prove that the fraction of galaxies dominated by a contribution of AGN origin is drastically reduced at faint mid-IR fluxes (e.g. Brand et al. 2006; Weedman et al. 2006; Treister et al. 2006; Magliocchetti et al. 2007). Unfortunately, the exact proportion of AGN vs starforming dominated galaxies is still undetermined. This separation is further complicated by the existence of a noticeable number of mixed systems where both star formation and AGN activity significantly contribute to the IR emission. For instance, Daddi et al. (2007a) find that about 20\\% of 24$\\mu$m-selected galaxies in the GOODS sample show a mid-IR excess which is not possible to reconcile with pure star-forming activity. Such a fraction increases to $\\sim 50-60$\\% at the highest (stellar) masses probed by their study. Clearly, understanding how the mid-IR sources divide between starbursts, AGN and composite systems is now the next essential step in order understand the relationships amongst the formation and evolution of stars, galaxies and massive black holes powering AGNs within dusty environments and more generally within massive systems observed at the peak of their activity. This paper approaches the study of the population of optically faint luminous infrared galaxies from the point of view of their multifrequency radio emission. Diagnostics based on the radio signal steming from these sources can in fact provide precious information on the process(es) which are actively taking place within such systems. Enhanced radio activity can stem from supernova remnants associated with regions which are vigorously forming stars, or originate from nuclear activity (AGN-dominated sources). These two processes determine a rather different spectral behaviour at radio wavelengths: star-forming systems are in general characterized by radio spectra which feature power-law shapes with slope (hereafter called {\\it radio spectral index} $\\alpha$, defined as $F\\propto \\nu^{-\\alpha}$, with $F$ radio flux and $\\nu$ radio frequency) of the order of $\\sim$0.7-0.8, while typical radio-loud quasars exhibit values for $\\alpha$ between 0 and 0.5 even though, especially at high redshifts, there is a non negligible population of radio-loud sources with very high, $\\alpha>1$, values (see e.g. De Breuck et al. 2000). Radio counterparts to the 510 optically faint mid-IR selected sources drawn from the whole First Look Survey sample (Fadda et al. 2006) have been searched in the overlapping region between MIPS and IRAC observations (Magliocchetti et al. 2007). Despite not having direct redshift estimates except for a handful of cases (e.g. Weedman et al. 2006; Yan et al. 2005; 2007; Martinez-Sansigre et al. 2006a), mid-IR photometry indicates that the overwhelming majority of such sources reside at redshifts $1.7\\simlt z\\simlt 2.5$ (\\S~4; see also Brand et al. 2007; Houck et al. 2005; Weedman et al. 2006a).\\\\ The First Look Survey region provides an excellent laboratory for investigations of the radio properties of dusty galaxies set at redshifts $z\\sim 2$ as its area has been observed at a number of radio frequencies down to very low flux densities (Condon et al. 2003; Morganti et al. 2004; Garn et al. 2007), therefore maximizing our chances of finding radio emitting galaxies. The layout of the paper is as follows. In \\S2 we introduce the parent (mid-IR and radio) catalogues, while in \\S3 we present the matching procedure leading to the sample of optically obscured {\\it Spitzer}-selected sources with a radio counterpart at 1.4~GHz and/or 610~MHz. \\S4 uses IRAC photometry to provide some information on the typology of such objects (i.e. whether mainly powered by an obscured AGN or by a starburst) and also -- where possible -- to assign them to a redshift interval. \\S5 presents the results on the radio number counts both at 1.4~GHz and 610~MHz (\\S5.1) and on the relationship between radio and 24$\\mu$m emission (\\S5.2). \\S6 discusses our findings on the 1.4~GHz vs 610~MHz radio spectral indices for the objects in our sample, while \\S7 summarizes our conclusions. ", "conclusions": "This paper has presented an analysis of the radio properties of a subsample of optically faint ($R\\simgt 25.5$), 24$\\mu$m-selected galaxies observed by {\\it Spitzer} in the FLS (Magliocchetti et al. 2007). These objects have been cross-correlated with a number of radio catalogues which cover the same region of the sky, namely that of Condon et al. (2003) which probes 1.4~GHz fluxes brighter than $\\sim 100 \\mu$Jy, that of Garn et al. (2007) -- which probes 610~MHz fluxes brighter than $\\sim 200 \\mu$Jy -- and, on a smaller portion of the sky, that of Morganti et al. (2004) which reaches 1.4~GHz fluxes as faint as $\\sim 40\\mu$Jy. 70 optically faint {\\it Spitzer} sources have been identified in the Condon et al. (2003) catalogue, 33 in the Morganti et al. (2004) dataset, while 52 are found in the survey performed by Garn et al. (2007). After performing a number of corrections to account for multiple identifications, sources erroneously split in the original {\\it Spitzer} catalogue into different components and mid-IR objects with real radio counterparts at one of the two radio frequencies which were further away than the allowed ($10^{\\prime\\prime}$) matching radius, we end up with a sample of 96 radio-identified, optically faint, mid-IR emitting sources, 45 of which have an identification at both 1.4~GHz and 610~MHz. The fraction of radio identifications is a strong function of 24$\\mu$m flux: almost all sources brighter than $F_{24\\mu\\rm m}\\sim 2$~mJy are endowed with a radio flux at both 1.4~GHz and 610~MHz, while this fraction drastically decreases by lowering the flux level. IRAC photometry for all those sources which also have detected fluxes in at least one of the four 8$\\mu$m, 5.8$\\mu$m, 4.5$\\mu$m and 3.6$\\mu$m channels (64 out of 96), allows to classify them into two categories: obscured AGN (45 sources) and systems mainly powered by starformation activity, (SB, 13 objects). We also find five low-z (i.e. $z\\simlt 0.5$ m51-type) interlopers, while the remaining 33 sources are unclassified. Furthermore, with the help of IRAC photometry it was possible to assign broad redshift intervals to all those sources (mostly AGN) which presented in the lowest 3.6$\\mu$m and $4.5\\mu$m wavelength channels of IRAC a 'bump' compatible with being produced by an evolved (old) stellar population. The majority ($\\sim 66$\\%) of these galaxies reside at redshifts $z\\simgt 1$, in agreement with other studies mainly based on mid-IR and near-IR spectroscopy of optically faint, 24$\\mu$m-selected galaxies (i.e. Weedman et al. 2006; Yan et al. 2005; 2007; Brand et al. 2007). We stress that this inferred fraction can only be considered as a lower limit to the real portion of faint {\\it Spitzer} sources set at high redshifts. In fact, because of their characteristic spectral properties in the mid-IR regime, we expect the majority of star forming systems (too faint at the IRAC frequencies to have measurable 3.6$\\mu$m and/or 4.5$\\mu$m fluxes) to be indeed located in the $z\\sim [1.6-2.7]$ redshift range. A small fraction of objects in our sample present radio morphologies such as jets and/or lobes compatible with them being identified as radio-loud AGNs. Interestingly enough, we find that for most of these few extended radio sources the mid-IR emission is associated to such peripheral regions rather than steming from the centre of radio activity, generally coinciding with the location of the AGN. However, the majority of the objects in our sample present unresolved radio images. A compared analysis of the radio number counts for optically obscured {\\it Spitzer} sources indicates that the $\\Delta N/\\Delta F$ as estimated at 1.4~GHz can only be reconciled with what found at 610~MHz if the population under investigation is endowed with an average value for the radio spectral index (defined as $F_R\\propto \\nu_R^{-\\alpha}$, where $F_R$ is the radio flux and $\\nu_R$ the generic radio frequency) $\\left<\\alpha\\right>\\simgt 1$. Classical, 'flat spectrum' radio sources can confidently be excluded as the typical objects constituting our sample. Furthermore, we have found that the radio number counts of sources classified as AGN and of those identified as starburst galaxies are quite similar, evidence which suggests that radio emission in $z\\sim 2$ {\\it Spitzer} galaxies originates from similar process(es), despite of the different mid-IR emission. Direct investigations of the relation between 24$\\mu$m and 1.4~GHz fluxes show that the overwhelming majority of those galaxies not excluded from our analysis because morphologically classified as radio-loud AGN follow the relationship identified for $0\\simlt z\\simlt 1$ star-forming objects by Appleton et al. (2004), although with a large scatter. This happens regardless of whether the galaxy has been classified as an AGN or a starforming system on the basis of its mid-IR colours. The distribution of 24$\\mu$m vs 610~MHz fluxes is instead found to be more scattered. The majority of these radio-identified objects (26, a figure which rises to 34 if one also includes sources with estimated lower limits on $\\alpha$) present very steep, $\\alpha> 1$ radio spectral indices, some galaxies being endowed with $\\alpha$'s as high as 2.5. This excess of galaxies with large $\\alpha$ values is statistically significant as it corresponds to 60 per cent of our sample, to be compared with the 43 per cent found by considering {\\it the whole population} of faint radio objects with an identification at both 610~MHz and 1.4~GHz.\\\\ Such very high figures for $\\alpha$ would identify the corresponding sources as Ultra Steep Spectrum galaxies, generally high redshift radio-loud AGN. However this is in striking disagreement with what found for the relation between 24$\\mu$m and 1.4~GHz fluxes for the very same objects, relation which would explain their (1.4~GHz) radio emission as mainly due to processes connected with star forming activity. A natural explanation to the above issues could be found by assuming that AGN and star-formation activity are concomitant in the majority of $z\\sim 2$, {\\it Spitzer} sources, at least in those which present enhanced radio emission. The radio signal steming from these systems would then simply be the combination of two components: a shallower one -- dominating the spectrum at 1.4~GHz -- due to processes connected with star formation, and a steeper one -- being responsible for most of the 610~MHz signal -- connected with AGN activity. This framework could then also explain why the 1.4~GHz emission in our sources follows that of star forming systems, while the same does not seem to happen in the case of 610~MHz fluxes. Various explanations can be found in the literature to account for the presence of USS: from slow adiabatic expansion losses in high-density environments (e.g. Klamer et al. 2006) to the scattering between CMB photons and relativistic electrons at $z\\sim 2$ (e.g. Martinez-Sansigre et al. 2006). However, the results presented in this work might provide an alternative scenario. In fact, they suggest that high values for $\\alpha$ might be due to the concomitant presence within the same systems of an AGN and of a star forming region: the AGN expansion would then be halted by the encounter with the cooler/denser sites in which star formation takes place. This would determine the 'strangling' of the AGN, causing its radio spectrum at low radio frequencies to steepen. Clearly, more theoretical work is needed in order to quantify the above issue and we are planning to present it in a forthcoming paper. For the time being, we note that intense star forming activity within a high redshift galaxy host of a USS has been recently reported by Hatch et al. (2007). From a more observational point of view, the 'ultimate truth' on these sources could only come from very high resolution (and, given their faintness, sensitivity) measurements, capable to clearly disentangle the emission related to the AGN to that associated to star forming regions. The advent of instruments such as ALMA will then provide the answers we need." }, "0710/0710.2452_arXiv.txt": { "abstract": "We present highlights and an overview of 20 {\\em FUSE} and {\\em HST}/STIS observations of the bright symbiotic binary \\object{EG And}. The main motivation behind this work is to obtain spatially-resolved information on an evolved giant star in order to understand the mass-loss processes at work in these objects. The system consists of a low-luminosity white dwarf and a mass-losing, non-dusty M2 giant. The ultraviolet observations follow the white dwarf continuum through periodic gradual occultations by the wind and chromosphere of the giant, providing a unique diagnosis of the circumstellar gas in absorption. Unocculted spectra display high ionization features, such as the \\ion{O}{6} resonance doublet which is present as a variable ({\\em hourly} time-scales), broad wind profile, which diagnose the hot gas close to the dwarf component. Spectra observed at stages of partial occultation display a host of low-ionization, narrow, absorption lines, with transitions observed from lower energy levels up to $\\sim$5 eV above ground. This absorption is due to chromospheric/wind material, with most lines due to transitions of \\ion{Si}{2}, \\ion{P}{2}, \\ion{N}{1}, \\ion{Fe}{2} and \\ion{Ni}{2}, as well as heavily damped \\ion{H}{1} Lyman series features. No molecular features are observed in the wind acceleration region despite the sensitivity of {\\em FUSE} to H$_2$. From analysis of the ultraviolet dataset, as well as optical data, we find that the dwarf radiation does not dominate the wind acceleration region of the giant, and that observed thermal and dynamic wind properties are most likely representative of isolated red giants. ", "introduction": " ", "conclusions": "" }, "0710/0710.1876_arXiv.txt": { "abstract": "{During the period 1966.5--2006.2 the 15GHz and 8GHz lightcurves of 3C454.3 (z=0.859) show a quasi-periodicity of $\\sim$12.8 yr ($\\sim$6.9 yr in the rest frame of the source) with a double-bump structure. This periodic behaviour is interpreted in terms of a rotating double-jet model in which the two jets are created from the black holes in a binary system and rotate with the period of the orbital motion. The periodic variations in the radio fluxes of 3C454.3 are suggested to be mainly due to the lighthouse effects (or the variation in Doppler boosting) of the precessing jets which are caused by the orbital motion. In addition, variations in the mass-flow rates accreting onto the black holes may be also involved.\\\\ ", "introduction": "In recent years many works have been devoted to the study of the periodic properties discovered in extragalactic sources, especially in Blazars. These include the periodic (or quasi-periodic) variations of flux density (or luminosity) in optical bands (for example, for OJ287: Sillanp{\\\"a}{\\\"a} et al. \\cite{Si88}, Lehto \\& Valtonen \\cite{Le96}, Valtonen et al. \\cite{Va99}, Villata et al. \\cite{Vi98}, Valtaoja et al. \\cite{Va00}, Valtonen \\cite{Va06a}, \\cite{Va06b}; AO\\,0235+16: Raiteri et al. \\cite{Ra01}); periodic variations of flux density in radio bands (for example, for AO0235+16: Raiteri et al. \\cite{Ra01}, 3C454.3: Ciaramella et al. \\cite{Ci04}, Kudryavtseva \\& Pyatunina \\cite{Ku06}); the periodic changes of the position angle of the ejection of superluminal knots (for example, Qian et al. \\cite{Qi01}, Britzen et al. \\cite{Br01}, Abraham \\& Carrara \\cite{Ab00}, Lobanov \\& Roland \\cite{Lo05}); periodic changes of the trajectory of superluminal knots (for example, for 3C345: Qian et al. \\cite{Qi91}, Lobanov \\& Roland \\cite{Lo05}, Klare et al. \\cite{Kl05}), etc. Up to now, the most reliable periodic phenomenon observed in blazars may be the 12 year quasi-periodic optical variations of the BL Lac object OJ287 which has been convincingly determined in the more than 100 year long records of optical observations since 1890.\\\\ In the interpretations of these periodic phenomena various binary balck hole models have been invoked. For the periodic optical variations in OJ287 all the proposed models invoke a binary system with two supermassive black holes. But the mechanisms for the production of the optical outbursts can be divided into two classes: accretion models and lighthouse models. \\begin{itemize} \\item Accretion models\\\\ These models (Lehto \\& Valtonen \\cite{Le96}, Valtonen \\cite{Va06a}, \\cite{Va06b}) propose that the optical periodicity is caused by a precessing binary black hole system of two supermassive black holes with a large eccentric orbit. The observed periodic 'superflares' with double peaks are suggested to be due to the crossings of the secondary black hole into the accretion disk of the primary black hole during the pericenter passage. Thus in this class of models the optical emission is thermal (bremsstrahlung) and the observed flux increase reflects the enhanced accretion related to the disk crossings and the general enhancement of the accretion rate during the pericenter passage. Doppler boosting is not taken into account. Detailed models for a binary black hole system have been proposed to fit the past optical records for OJ287 and predict the future optical events. If the timing of the optical activity (superflares) in OJ287 can be accurately predicted, then it is possible to calculate or determine the parameters of the binary black hole system (such as the masses of the black holes, the orbital parameters etc.). \\item Lighthouse models\\\\ The periodic optical flares observed in OJ287 are suggested to be caused by the change of the orientation of the realtivistically moving emission regions with respect to the line of sight, resulting in an increase of the Doppler boosting factor (Camenzind \\& Krockenberg \\cite{Ca92}, Katz \\cite{Ka97}, Villata et al. et al. \\cite{Vi98}, Qian et al. \\cite{Qi01}) and the apparent optical flux density (${\\propto}{{\\delta}^3}$). Katz (\\cite{Ka97}) proposed that in a bianry black hole system the companion exerts a torque on the accretion disk of the primary black hole and results in the precessing of the relativistic jet from the primary sweeping across the line of sight and the periodic variation of the Doppler factor and thus the apparent flux density. Villata et al. (\\cite{Vi98}) suggests that the two black holes of the binary system both create relativistic jets which are bent significantly in different directions. In the course of the binary's orbit motion, the directions of the bent parts of the jets from the two black holes rotate with the orbital period, resulting in periodic double-peak flares. In this class of models only the variation of the Doppler boosting factor plays the role, not considering any change in accretion in the central disk-hole systems. \\item Alternative models\\\\ Valtaoja et al. (\\cite{Va00}) have argued against both the accretion models and lighthouse models for OJ287. They found that in OJ287 the first optical flare of the double-structure is thermal (non-polarized) and lacks radio counterpart, but the second one is (polarized) synchrotron emission, having a radio counterpart. They proposed an alternative model in which the first optical flare is caused by the disk-crossing during the pericenter passage, having a regular period. And the second one occurs about one year later in the relativistic jet with an association with a radio outburst. For checking this model optical polarization measurements and VLBI studies of the optical-radio association are significant. \\end{itemize} In this paper we will discuss the possible existence of a $\\sim$13 year periodicity in the radio lightcurve of the optically violently variable (OVV) quasar 3C454.3 at 8GHz and 14.5GHz in which two broad peaks (or bumps) are observed during one period. We propose that the periodicity can be interpreted in the frame of a binary black hole model in which two jets from the two black holes rotate with the period ($\\sim$13 year) of the orbital motion. Model-fits to the lightcurves are given. It is shown that both the periodic Doppler boosting effect (lighthouse effect) and the variability of the accretion activity (i.e. the mass-energy inflow into the jets) should be taken into account in order to fully explain the lightcurves. We also discuss some alternative models and future observations for further studying the periodic phenomenon in 3C454.3. ", "conclusions": "\\begin{itemize} \\item (1) the supposed model is not unique. In addition to the double-jet models proposed in this paper, models with a single jet are also plausible. But in this case, the two emitting components could be assumed to be situated at different positio ns in the curved jet and their rotation would produce the Doppler boosting profiles. In this paper we do not invoke a specific model for a binary black hole system. \\item (2) In our model-fittings, multi-parameters can be chosen. When parameters were chosen we consider mainly three ingredients: (a) the width of the two radio bumps; (2) the difference of Doppler factor at flaring and quiescent epochs should not be too large in order to avoid a huge difference in timescale of varibility at different epochs (Valtaoja et al. \\cite{Va00}); \\item (3) We should consider both effects due to Doppler boosting and the energy transfer into the jets (or accretion rates). In our models, the Doppler boosting determine the 'profiles' of the possible activity in the radio and optical bands. If the intrinsic (in the rest frame of the flows) strengths are stable, then the apparent activity follows the pattern of the Doppler boosting. This seems correct for the observed radio variations, for example during 1978--1995. \\item (4) As for the optical variations, they do not follow the Doppler boosting profiles derived from our models. This could imply we should adopt different parameters for the optical variations, i.e., the optically emitting regions could have orientations different from the radio regions. Moreover, optical flux variations could be largely determined by the variation in accretion rate and their much shorter timescales may be due to mechanisms for acceleartion of electrons and radiative losses. The optical peak at 2005.4 and radio peak at 2006.2, which both deviate the Doppler boosting pattern maximum, further indicate the necessity to consider the effects both from Doppler boosting, accretion and transfer of enery into the jets. \\item (5) The radio periodic behaviour still needs to be tested, because the 2005 optical flare has a quite peculiar mm-radio outbursts, which have much narrower profiles than before, very different from the 1981 and 1994 radio flare's profiles. This may imply that the radio periodic behaviour could be a short-term phenomenon. The optical-radio relation is still important to be tested and thus could obtain some information about the combination of the black hole disk system and the Doppler boosting effect. \\item (6) Relationship between mm- and cm-outburtsts needs to be further studied in order to study the evolution of the outbursts. \\item (7) At present, we cannot predict whether the periodic behaviour in 3C454.3 in radio and optical bands can hold long, and this should be observationally tested during the next period (2007--2020). Both optical and radio monitoring (intensity and polarization) are required. \\end{itemize}" }, "0710/0710.3042_arXiv.txt": { "abstract": "{We show that future Ultra-High Energy Cosmic Ray samples should be able to distinguish whether the sources of UHECRs are hosted by galaxy clusters or ordinary galaxies, or whether the sources are uncorrelated with the large-scale structure of the universe. Moreover, this is true independently of arrival direction uncertainty due to magnetic deflection or measurement error. The reason for this is the simple property that the strength of large-scale clustering for extragalactic sources depends on their mass, with more massive objects, such as galaxy clusters, clustering more strongly than lower mass objects, such as ordinary galaxies. } \\begin{document} ", "introduction": "\\label{sec:intro} Identifying the sources of ultrahigh energy cosmic rays (UHECRs, here $E>10^{19}$eV $\\equiv 10$ EeV) is complicated by the deflection they presumably experience in Galactic and extragalactic magnetic fields, as well as their relatively poor arrival direction determinations, typically $\\sim 1^{\\circ}$. Arrival directions of most UHECRs are thus not known well enough to match their positions with specific astrophysical objects. However, there is also useful information in the clustering of UHECRs on large scales, where $\\sim$ few degree uncertainties in position become unimportant. The clustering of galaxies in the universe is typically quantified by the two-point correlation function or its analog in Fourier space, the power spectrum. The two-point correlation function $\\xi(r)$ of any class of objects (e.g., galaxies of a certain luminosity or color) is defined as the excess number of pairs of such objects at physical separation $r$ over that expected for a random (Poisson) distribution. In Cold Dark Matter models, the large-scale amplitude of $\\xi(r)$ (usually referred to as the bias) of a population of objects depends only on their mass, with more massive objects, such as clusters of galaxies, clustering more strongly than less massive objects, such as ordinary galaxies \\cite{mo_white_96,sheth_tormen_99,berlind_etal_06b}. The large-scale bias of a UHECR sample is therefore a robust and informative measure of the clustering properties of the source. We cannot measure physical separations for pairs involving UHECRs because they do not have measured redshifts. However, we can measure the angular correlation function $\\omega(\\theta)$. As is the case for $\\xi(r)$, the large-scale amplitude of $\\omega(\\theta)$ for a UHECR sample depends on the nature of the astrophysical source. However, it also depends on the depth of the sample because deeper samples mix more physically uncorrelated pairs and thus show weaker angular clustering. In order to access the information in the large-scale angular clustering of UHECRs, we must therefore know the depth of our UHECR sample. In this paper, we demonstrate what can be learned from the large-scale angular clustering of UHECRs, we estimate what kind of sample is needed to do this analysis, and we show how to deal with the unknown depth of a UHECR sample, using the GZK effect. ", "conclusions": "" }, "0710/0710.4986_arXiv.txt": { "abstract": "We calculate axisymmetric toroidal modes of magnetized neutron stars with a solid crust. We assume the interior of the star is threaded by a poloidal magnetic field that is continuous at the surface with the outside dipole field whose strength $B_p$ at the magnetic pole is $B_p\\sim10^{16}$G. Since separation of variables is not possible for oscillations of magnetized stars, we employ finite series expansions of the perturbations using spherical harmonic functions to represent the angular dependence of the oscillation modes. For $B_p\\sim 10^{16}$G, we find distinct mode sequences, in each of which the oscillation frequency of the toroidal mode slowly increases as the number of radial nodes of the eigenfunction increases. The frequency spectrum of the toroidal modes for $B_p\\sim10^{16}$G is largely different from that of the crustal toroidal modes of the non-magnetized model, although the frequency ranges are overlapped each other. This suggests that an interpretation of the observed QPOs based on the magnetic toroidal modes may be possible if the field strength of the star is as strong as $B_p\\sim10^{16}$G. ", "introduction": "Recent discovery of quasi-periodic oscillations (QPOs) of magnetar candidates is one of the observational manifestations of global oscillations of neutron stars. Israel et al (2005) detected QPOs of frequencies $\\sim18$, $\\sim 30$ and $\\sim$92.5Hz in the tail of the SGR 1806-20 hyperflare observed December 2004, and suggested that the 30Hz and 92.5Hz QPOs could be caused by seismic vibrations of the neutron star crust (see, e.g., Duncan 1998). Later on, in the hyperflare of SGR 1900+14 detected August 1998, Strohmayer \\& Watts (2005) found QPOs of frequencies 28, 53.5, 84, and 155 Hz, and claimed that the QPOs could be identified with the low $l$ fundamental toroidal torsional modes of the solid crust of the neutron star. These recent discoveries of QPOs in the giant flares of Soft Gamma-Ray Repeaters SGR 1806-20 (Israel et al 2005, Watts \\& Strohmayer 2006, Strohmayer \\& Watts 2006) and SGR 1900+14 (Strohmayer \\& Watts 2005) have made promising asteroseismology for magnetars, neutron stars with an extremely strong magnetic field (see, e.g., Woods \\& Thompson 2006 for a review of SGRs). It is currently common to identify these QPOs with seismic vibrations caused by crustal toroidal modes of the neutron stars, since the frequency range of the modes overlap that of the observed QPOs and from the energetics point of view the crustal toroidal modes would be most easily excited to observable amplitudes by spending a least amount of available energies, for example, those released in magnetic field restructuring (e.g., Duncan 1998). Although the interpretation based on crustal torsional modes looks promising, we need detailed theoretical analyses of oscillations of magnetized neutron stars so that we could get information of physical conditions of the stars through the confrontation between theoretical modelings and observations. This is particularly true for high frequency QPOs (e.g., 625Hz QPO, Watts \\& Strohmayer 2006; 1835Hz QPO and less significant QPOs at 720 and 2384 Hz in SGR 1806-20, Strohmayer \\& Watts 2006), since there exist classes of modes other than the crustal toroidal modes that can generate the frequencies observed. The presence of a magnetic field makes it possible for toroidal modes to exist in a fluid star even without rotation as does the presence of the shear modulus in the solid crust. In this paper we are interested in axisymmetric toroidal modes since axisymmetirc toroidal and spheroidal modes are decoupled for a poloidal field when the star is non-rotating. For non-axisymmetric modes, the toroidal and spheroidal components are coupled even without rotation and hence the modal analyses of magnetized stars would be much more complicated. Theoretical calculations of toroidal modes of strongly magnetized neutron stars have been carried out by several authors, including Piro (2005), Glampedakis et al (2006), Sotani et al (2006, 2007), and Lee (2007). The analyses by Piro (2005), Glampedakis et al (2006), and Lee (2007) assume Newtonian gravity, while those by Sotani et al (2006, 2007) use general relativistic formulation. Although the studies by Piro (2006) and Lee (2007) ignore the effects of magnetic fields in the fluid core, those by Glampedakis et al (2006) and Sotani et al (2006, 2007) consider magnetic waves propagating in the fluid core, assuming the core is threaded by a magnetic field of substantial strength. Besides the differences mentioned above, most of the authors except Lee (2007) represent the angular dependence of the oscillations by a single spherical harmonic function $Y_l^m(\\theta,\\phi)$. Since the shear modulus in the crust dominates the magnetic pressure in most parts of the crustal regions for a dipole field of strength $B_p\\ltsim 10^{15}$G, this treatment may be justified so long as the crustal toroidal modes are well decoupled from the fluid core. But, if the torsional waves in the crust are strongly coupled with magnetic waves in the core, the treatment may not be justified because the angular dependence of the magnetic waves in the fluid core cannot be correctly represented by a single spherical harmonics. For example, the analysis by Reese, Fincon, \\& Rieutord (2004) of toroidal modes in a fluid shell have employed finite series expansions of long length for the perturbations. In this paper, using the method of series expansions of perturbations we calculate toroidal modes of a strongly magnetized neutron star having a fluid core and a solid crust, where the entire interior is assumed to be threaded by a poloidal magnetic field. We employ two different sets of oscillation equations, one for fluid regions and the other for the solid crust, and solutions in the solid and fluid regions are matched at the interfaces between them to obtain an entire solution of a mode. The method of calculation we employ is presented in \\S 2, and the numerical results are given in \\S 3, and conclusions are in \\S 4. ", "conclusions": "In this paper, we have calculated toroidal modes of magnetized stars with a solid crust, where the entire interior of the star is assumed to be threaded by a poloidal magnetic field that is continuous at the stellar surface to the outside dipole field. We find distinct mode sequences of the toroidal modes, in each of which the mode frequency remains rather constant and only slowly increases as the radial order of the modes increases. In the presence of a solid crust, the frequency separation between the lowest and second lowest frequency mode sequences is approximately given by $\\Delta\\omega$, but that between the second and third lowest frequency mode sequences by $\\Delta\\omega/2$, where the frequency separation $\\Delta\\omega$ is roughly proportional to the field strength $B_p$. This frequency pattern of the low frequency sequences is different from that found for the model without the crust, for which the frequency separation between the sequential sequences of low frequency modes is given by $\\Delta\\omega/2$. We also find that for the equation of state we use, $\\Delta\\omega$ is larger for smaller $M$. The eigenfunction $\\xi_\\phi$ of the modes belonging to the lowest frequency sequence have much larger amplitudes than $B^\\prime_\\phi$ and can penetrate into the solid crust. On the other hand, $\\xi_\\phi$ of the modes belonging to the higher frequency sequences is much smaller than $B^\\prime_\\phi$, which is well confined into the fluid core and does not have any substantial amplitudes in the solid crust. The frequency ranges of the toroidal modes we find for the magnetized neutron star with $B_p\\sim10^{16}$G overlap the QPO frequencies found for the magnetar candidates, SGR 1806-20 and SGR 1900+14. This suggests that we may interpret the observed QPOs based on the magnetic toroidal modes, and that detailed comparisons between observed frequency spectra and theoretical calculations make it relevant to infer physical parameters of the magnetar candidates, such as the equation of state and the strength of the magnetic field. But, we think it worth pointing out that except for the modes belonging to the lowest frequency sequence, the magnetic perturbations, which have much larger amplitude than $\\xi_\\phi$, are well confined in the fluid core and do not have substantial amplitudes in the solid crust, which make it difficult for the modes to be directly observable. If the magnetar candidates do no have a crust, the problem of observability could be avoided. In this case, however, we have to use more realistic surface boundary conditions than $\\pmb{B}^\\prime=0$ used in the present calculations. Note that, although the existence of a solid crust affects the frequency pattern of the low frequency mode sequences, the frequency range of the magnetic modes itself is not very much dependent on the presence or absence of a solid crust unless the crust is extremely thick. We have tried to find toroidal modes well confined in the solid crust or core toroidal modes that are in resonance with the crustal toroidal modes, but failed. We also find it extremely difficult to identify distinct toroidal mode sequences for magnetic fields weaker than $B_p\\ltsim10^{15}$G when a solid crust is included in the models. It is therefore not clear whether the toroidal mode sequences of the kind we find in the present paper can survive also for weakly magnetized neutron stars with a solid crust. If the field strength is much weaker than $B_p\\sim10^{15}$G, the magnetic fields may have only minor effects on the crustal toroidal modes (e.g., Lee 2007), and it will be justified to use frequency spectra of the crust modes theoretically obtained in the weak field limit to interpret observed QPOs. As briefly noted in the last section, it is not theoretically clear how the frequency spectra of the toroidal modes of magnetized stars should look like, and how the toroidal modes behave in the limit of $n\\ge k\\rightarrow\\infty$. We may even speculate that the frequency spectra we obtained reflect the existence of continuous spectra (see, e.g., Goedbloed \\& Poedts 2004, see also Levin 2007), but the detailed numerical analysis of continuous frequency spectra is beyond the scope of this paper. It will be worthwhile to examine the effects of an interior toroidal field on magnetic modes. It is also needed to extend the present analysis to a general relativistic formulation (e.g., Sotani et al 2006, 2007)." }, "0710/0710.1271_arXiv.txt": { "abstract": "We present results of 3-neutrino flavor evolution simulations for the neutronization burst from an O-Ne-Mg core-collapse supernova. We find that nonlinear neutrino self-coupling engineers a single spectral feature of stepwise conversion in the inverted neutrino mass hierarchy case and in the normal mass hierarchy case, a superposition of two such features corresponding to the vacuum neutrino mass-squared differences associated with solar and atmospheric neutrino oscillations. These neutrino spectral features offer a unique potential probe of the conditions in the supernova environment and may allow us to distinguish between O-Ne-Mg and Fe core-collapse supernovae. ", "introduction": " ", "conclusions": "" }, "0710/0710.3104_arXiv.txt": { "abstract": "The Milky Way is the only galaxy for which we can resolve individual stars at all evolutionary phases, from the Galactic center to the outskirt. The last decade, thanks to the advent of near IR detectors and 8 meter class telescopes, has seen a great progress in the understanding of the Milky Way central region: the bulge. Here we review the most recent results regarding the bulge structure, age, kinematics and chemical composition. These results have profound implications for the formation and evolution of the Milky Way and of galaxies in general. This paper provides a summary on our current understanding of the Milky Way bulge, intended mainly for workers on other fields. ", "introduction": "How did the Milky Way form? Decades ago, the Galactic formation scenarios focused on the disk and the halo populations, because these were the components that astronomers knew something about. For example, in the classic works of Eggen et al. (1962) or Searle \\& Zinn (1978) the bulge was not mentioned. All-sky optical maps (Fig.~\\ref{maps}) did not show a clear, separate component in the inner Milky Way. Yet upon looking at the new DIRBE or 2MASS near-infrared maps of the whole sky, it is evident that the Milky Way is a spiral galaxy with a peanut-shape bulge. The simultaneous observation that bulge stars were mainly old made it clear that to answer this question the attention had to shift towards what seems to be the first massive component to be formed in the Milky Way. This is a Copernican-like revolution on Galactic scales, and it is happening now! We see our Galaxy as if it were an external galaxy for the first time, and are fortunate to have such new perspective, and also to be able to provide specific answers to the many questions regarding its formation. \\begin{figure} \\includegraphics[height=5.5in,width=5.3in,angle=00]{minniti_fig1.ps} \\caption{ Top: Near-IR COBE-DIRBE map of the whole sky (Dwek et al. 1995). Bottom: Optical map for comparison (Copyright Axel Mallenhoff 2001). These sky maps illustrate why astronomers did not realize before the importance of the bulge.} \\label{maps} \\end{figure} Note that we have only one object to study: our Galaxy. This is a fundamental limitation because we have to get the answers right. The advantage in the case of our Galaxy, of course, is that we can study the subject in detail: in no other galaxies the fundamental problems of Astronomy can be surpassed. Basically, these fundamental problems are: (1) The distance problem: we can resolve the components of the Milky Way into stars, and study them in situ, obtaining 3-D positions and motions, and detailed chemical compositions of individual stars; and (2) The timescale problem: we see an instant snapshot, and must infer histories from that. In our galaxy we can separate and date the components, using the main-sequence turn-offs of different stellar populations and clusters. In spite of a small community working on the Galactic bulge, a revolution has occurred in this field in the last decade, with great progress in comparison with other fields of Astronomy. For the first time, we have the answers to important questions such as: {\\bf 1.} Is the bulge a different Galactic component? {\\bf 2.} When did the Galactic bulge form? {\\bf 3.} How did it form? {\\bf 4.} Is there a radial gradient in the bulge? {\\bf 5.} Are there globular clusters associated with the bulge? {\\bf 6.} Are there planets in the bulge? The 8m class telescopes were built last decade, aiming to answer these questions. The proposed options were endless (bulge formation from secular evolution of the disk, from the halo, in a single burst, in several episodes, by slow accretion of smaller sub-units, etc). The evidence and the answers described below serve as a basis for understanding more distant galaxies which cannot be studied (dissected) in similar detail. ", "conclusions": "For the first time, we have the answers to the following basic questions: {\\bf 1.} Is the bulge a different component? Yes, based on all the evidence available, the bulge is a distinct Galactic component, with different kinematics and compositions from the thin disk, the thick disk and the halo. {\\bf 2.} How did the Galactic bulge form? It formed on a short timescale ($\\sim 1$ Gyr), as demonstrated by the $\\alpha$-element enrichment. Despite the presence of the bar, models like a bulge formation via secular evolution of the disk can be firmly excluded. {\\bf 3.} What is the age of the bulge? The bulk of the stellar population is $\\sim 10$ Gyr old. However, there are traces of a small fractions of intermediate-age stars, and of metal-poor stars. The latter might well be the oldest population in the Galaxy. {\\bf 4.} Is there a gradient in the bulge? Yes, there is a stellar population gradient as shown for example by the CMDs. Now it is found that this gradient is mostly due to metallicity, which decreases along the Galactic minor axis. {\\bf 5.} Are there globular clusters associated with the bulge? Yes, there is a population of metal-rich globular clusters in the central regions that share the kinematics, spatial distribution, and composition of the bulge field stars. {\\bf 6.} Are there planets in the bulge? Even though this question seems to belong to another field, another of the recent advances was the discovery of planets in the bulge by HST. The available data suggests that giant planets are as numerous in the bulge as they are in the Solar neighborhood. These revolutionary advances that impact the whole of extragalactic Astronomy cannot be attributed to the success of a single group, but to the combined contributions of different teams. Where controversy was present before, today similar answers are given. Progress has occurred!" }, "0710/0710.3618_arXiv.txt": { "abstract": " ", "introduction": "In the course of carrying out a wide-field CCD imaging survey, two new methods for correlating the images to star catalogues have been developed, motivated by the need to efficiently handle the large number of stellar sources present on the images. Most previously published algorithms successfully cater for small lists ($\\le$ 50 stars), but do not scale well to wide-fields containing $10^3$ or more stellar sources. The problem of matching coordinate lists of point sources is a necessary prerequisite for deriving an astrometric plate solution. The objective is to match a subset of stars found on an image to their corresponding entries in a stellar catalogue in order to determine the transformation between detector coordinates and sky coordinates. The algorithm must handle translation and rotation, and small changes in scale caused by temperature related changes in focal length. In addition, it must cope with additional and missing stars. That is, the two lists may only partially overlap. The efficiency of the algorithm is of paramount concern, since it is embodied within the closed-loop pointing system of the telescope and therefore affects the duty-cycle time, and ultimately constrains the number of images that can be acquired each night. Surveys that require very high photometric precision typically seek to accurately align their fields on the same detector pixels each night to overcome residual flat-fielding errors \\citep{Everett}, and would benefit from the efficiency gains of a fast matching algorithm. Similarly, high cadence surveys, such as the Southern Sky Survey \\citep{Keller} could improve precision and reduce its duty-cycle by utilizing a fast closed-loop pointing algorithm. Moreover, real-time attitude adjustments on spacecraft might be possible with the aid of an efficient matching algorithm to analyze on-board star camera images \\citep[see for example][]{Fraser}. A number of algorithms have been proposed to solve this problem. \\citet{Groth} describes an algorithm that matches geometrically similar shapes (triangles) in the two lists. By limiting the number of triangles constructed, and by only matching those triangles whose ratio of longest to shortest side are within a defined limit, his matching phase has a computational complexity of $O(n^{4.5})$ where $n$ is the number of stars in each list. \\cite{Stetson} describes a very similar algorithm that he developed independently at around the same time. \\citet{Murtagh} reviews a number of approaches and proposes his own, based upon characterization of a set of coordinates couples, with matching based on the proximity of feature vectors in the two lists. His method's matching phase has a computational complexity of $O(n^2)$. Nevertheless, Groth's algorithm appears to be the most widely accepted, with the methods applied across disciplines. For example, \\citet{Arz} discuss its application to the problem of computer-aided identification of whale sharks, while \\citet{Mars}, building upon the work of \\citet{Groth}, describe an optimization to the voting phase of the algorithm, concluding that their method reduces the need for complicated filtering methods while successfully reducing the number of false matches. More recently, \\citet{PB} describe another variation of triangle matching, optimized to handle large lists of objects extracted from wide field images. Large fields contain thousands of stars and pose a severe test for matching algorithms, requiring efficient methods to accommodate the large number of point sources. The following sections discuss two new methods for pattern matching that have a matching phase with a complexity that is nearly $O(1)$, at the cost of a slight loss in generality. They are collectively referred to as \\emph{Optimistic Pattern Matching (OPM)} because they assume that (i) a good match is likely to be found, and (ii) the scale of the image is approximately known, thus permitting the use of an early exit strategy whereby only a small percentage of the candidate list is examined. By contrast, previous methods assumed an unknown scale which required the entire candidate list to be processed to determine the most likely match using a statistical approach. This required additional phases and complexity. In practice, an \\emph{a priori} knowledge of an instrument's focal length is common place, and the use of a more general algorithm that assumes it is unknown mandates strategies that unnecessarily degrade performance. Section 2 describes the algorithms in detail. $OPM_A$ is based upon a new definition of triangle space, while $OPM_B$ uses an alternative shape characterization method. Section 3 tests their performance using a large sample of survey images and compares them to earlier methods. Conclusions are summarized in Section 4. ", "conclusions": "Two new techniques for matching two-dimensional coordinate lists in nearly constant time have been presented. The matching phase of $OPM_B$ is nearly $O(1)$, being independent of list size. These algorithms have a significant performance advantage over previous techniques, at a slight loss in generality, caused by the requirement that the approximate focal length of the optical system is known \\emph{a priori}. This requirement permits the determination of the image scale from the physical dimensions of the detector, allowing $OPM$ algorithms to directly compare a subset of triangles (or shapes) to their counterparts derived from a reference catalogue, without having to process the entire set, as is the case when the scale is unknown. By employing early exit strategies, postponing work until absolutely necessary, testing candidates in the order most likely to yield success, and combining these with and an efficient mechanism for rejecting false positives, a highly efficient search, in nearly constant time is possible. Small uncertainties in the focal length, such as caused by temperature related changes, are accommodated by selecting an appropriate matching tolerance. The actual focal-length is determined and reported as part of the astrometric solution. The $OPM$ algorithms are particularly suited to processing large lists or in situations where pattern matching must be performed as quickly as possible. The performance of these algorithms makes it practical to search thousands of fields very quickly, if for example, the coordinates of the field center were unknown. Similarly, when only an approximate focal-length is known, it is perfectly reasonable to attempt to iteratively match the field using a range of focal-lengths." }, "0710/0710.2102_arXiv.txt": { "abstract": "We use hydrodynamical simulations of disk galaxies to study relations between star formation and properties of the molecular interstellar medium (ISM). We implement a model for the ISM that includes low-temperature ($T<10^{4}$~K) cooling, directly ties the star formation rate to the molecular gas density, and accounts for the destruction of $\\Hmol$ by an interstellar radiation field from young stars. We demonstrate that the ISM and star formation model simultaneously produces a spatially-resolved molecular-gas surface density Schmidt-Kennicutt relation of the form $\\SigmaSFR \\propto \\SigmaHmol^{\\nmol}$ with $\\nmol\\approx1.4$ independent of galaxy mass, and a total gas surface density -- star formation rate relation $\\SigmaSFR \\propto \\Sigmagas^{\\ntot}$ with a power-law index that steepens from $\\ntot\\sim2$ for large galaxies to $\\ntot\\gtrsim4$ for small dwarf galaxies. We show that deviations from the disk-averaged $\\SigmaSFR \\propto \\Sigmagas^{1.4}$ correlation determined by \\cite{kennicutt1998a} owe primarily to spatial trends in the molecular fraction $\\fHmol$ and may explain observed deviations from the global Schmidt-Kennicutt relation. In our model, such deviations occur in regions of the ISM where the fraction of gas mass in molecular form is declining or significantly less than unity. Long gas consumption time scales in low-mass and low surface brightness galaxies may owe to their small fractions of molecular gas rather than mediation by strong supernovae-driven winds. Our simulations also reproduce the observed relations between ISM pressure and molecular fraction and between star formation rate, gas surface density, and disk angular frequency. We show that the Toomre criterion that accounts for both gas and stellar densities correctly predicts the onset of star formation in our simulated disks. We examine the density and temperature distributions of the ISM in simulated galaxies and show that the density probability distribution function (PDF) generally exhibits a complicated structure with multiple peaks corresponding to different temperature phases of the gas. The overall density PDF can be well-modeled as a sum of lognormal PDFs corresponding to individual, approximately isothermal phases. We also present a simple method to mitigate numerical Jeans fragmentation of dense, cold gas in Smoothed Particle Hydrodynamics codes through the adoption of a density-dependent pressure floor. ", "introduction": "\\label{section:introduction} Galaxy formation presents some of the most important and challenging problems in modern astrophysics. A basic paradigm for the dissipational formation of galaxies from primordial fluctuations in the density field has been developed \\citep[e.g.][]{white1978a,blumenthal1984a,white1991a}, but many of the processes accompanying galaxy formation are still poorly understood. In particular, star formation shapes the observable properties of galaxies but involves a variety of complicated dynamical, thermal, radiative, and chemical processes on a wide range of scales \\citep[see][for a review]{mckee2007a}. Observed galaxies exhibit large-scale correlations between their global star formation rate (SFR) surface density $\\SigmaSFR$ and average gas surface density $\\Sigmagas$ \\citep{kennicutt1989a,kennicutt1998a}, and these global correlations serve as the basis for treatments of star formation in many models of galaxy formation. While such models have supplied important insights, detailed observations of galaxies have recently provided evidence that the molecular, rather than the total, gas surface density is the primary driver of global star formation in galaxies \\citep[e.g.,][]{wong2002a,boissier2003a,heyer2004a,boissier2007a,calzetti2007a,kennicutt2007a}. In this study, we adopt an approach in which empirical and theoretical knowledge of the star formation efficiency (SFE) in dense, molecular gas is used as the basis for a star formation model in hydrodynamical simulations of disk galaxy evolution. This approach requires modeling processes that shape properties of the dense phase of the interstellar medium (ISM) in galaxies. The purpose of this paper is to present such a model. Stellar populations in galaxies exhibit salient trends of colors and metallicities with galaxy luminosity \\citep[e.g.,][]{kauffmann2003c,blanton2005a,cooper2007a}. In the hierarchical structure formation scenario these trends should emerge through the processes of star formation and/or stellar feedback in the progenitors of present-day galaxies. Observationally, ample evidence suggests that the efficiency of the conversion of gas into stars depends strongly and non-monotonically on mass of the system. For example, the faint-end of the galaxy luminosity function has a shallow slope \\citep[$\\alphaL\\approx1.0-1.3$, e.g.,][]{blanton2001a,blanton2003a} compared to the steeper mass function of dark matter halos \\citep[$\\alphaDM\\approx2$, e.g.,][]{press1974a,sheth1999a}, indicating a decrease in SFE in low-mass galaxies. At the same time, the neutral hydrogen (HI) and baryonic mass functions may be steeper than the luminosity function \\citep[$\\alphaHI\\approx1.3-1.5$, e.g.,][]{rosenberg2002a,zwaan2003a}. The baryonic \\cite{tully1977a} relation is continuous down to extremely low-mass dwarf galaxies \\citep[e.g.,][]{mcgaugh2005a,geha2006a}, indicating that the fractional baryonic content of galaxies of different mass is similar. Hence, low-mass galaxies that are unaffected by environmental processes are gas-rich, yet often form stars inefficiently. While feedback processes from supernovae and AGN \\citep[e.g.,][]{brooks2007a,sijacki2007a}, or the efficiency of gas cooling and accretion \\citep{dekel2006a,dekel2008a}, may account for part of these trends, the SFE as a function of galaxy mass may also owe to intrinsic ISM processes \\citep[e.g.,][]{tassis2008a,kaufmann2007a}. To adequately explore the latter possibility, a realistic model for the conversion of gas into stars in galaxies is needed. Traditionally, star formation in numerical simulations of galaxy formation is based on the empirical Schmidt-Kennicutt (SK) relation \\citep{schmidt1959a,kennicutt1989a,kennicutt1998a}, in which star formation rate is a {\\it universal} power-law function of the total disk-averaged or global gas surface density: $\\SigmaSFR\\propto \\Sigmagas^{\\ntot}$ with $\\ntot\\approx 1.4$ describing the correlation for the entire population of normal and starburst galaxies. However, growing observational evidence indicates that this relation may not be universal on smaller scales within galaxies, especially at low surface densities. Estimates of the slope of the SK relation within individual galaxies exhibits significant variations. For example, while \\citet{schuster2007a} and \\citet{kennicutt2007a} find $\\ntot\\approx 1.4$ for the molecular-rich galaxy M51a, similar estimates in other large, nearby galaxies \\citep[including the Milky Way (MW);][]{misiriotis2006a} range from $\\ntot\\approx1.2$ to $\\ntot\\approx 3.5$ \\citep{wong2002a,boissier2003a} depending on dust-corrections and fitting methods. The disk-averaged total gas SK relation for normal (non-starburst) galaxies also has a comparably steep slope of $\\ntot\\approx2.4$, with significant scatter \\citep{kennicutt1998a}. While the variations in the $\\SigmaSFR-\\Sigmagas$ correlation may indicate systematic uncertainties in observational measurements, intrinsic variations or trends in galaxy properties may also induce differences between the global relation determined by \\cite{kennicutt1998a} and the $\\SigmaSFR-\\Sigmagas$ correlation in individual galaxies. Galaxies with low gas surface densities, like dwarfs or bulgeless spirals, display an even wider variation in their star formation relations. \\citet{heyer2004a} and \\citet{boissier2003a} show that in low-mass galaxies the SFR dependence on the total gas surface density exhibits a power-law slope $\\ntot\\approx 2-3$ that is considerably steeper than the global \\citet{kennicutt1998a} relation slope of $\\ntot\\approx 1.4$. Further, star formation in the low-surface density outskirts of galaxies also may not be universal. Average SFRs appear to drop rapidly at gas surface densities of $\\Sigmagas\\lesssim 5-10{\\rm\\, \\Msun\\,pc^{-2}}$ \\citep{hunter1998a,martin2001a}, indicating that star formation may be truncated or exhibit a steep dependence on the gas surface density. The existence of such threshold surface densities have been proposed on theoretical grounds \\citep{kennicutt1998a,schaye2004a}, although recent GALEX results using a UV indicator of star formation suggest that star formation may continue at even lower surface densities \\citep{boissier2007a}. Star formation rates probed by damped Lyman alpha absorption (DLA) systems also appear to lie below the \\cite{kennicutt1998a} relation, by an order of magnitude \\citep{wolfe2006a,wild2007a}, which may indicate that the relation between SFR and gas surface density in DLA systems differs from the local relation measured at high $\\Sigmagas$. In contrast, observations generally show that star formation in galaxies correlates strongly with {\\it molecular\\/} gas, especially with the highly dense gas traced by HCN emission \\citep{gao2004a,wu2005a}. The power-law index of the SK relation connecting the SFR to the surface density of molecular hydrogen consistently displays a value of $\\nmol\\approx1.4$ and exhibits considerably less galaxy-to-galaxy variation \\citep{wong2002a,murgia2002a,boissier2003a,heyer2004a,matthews2005a,leroy2005,leroy2006,gardan2007a}. Molecular gas, in turn, is expected to form in the high-pressure regions of the ISM \\citep{elmegreen1993a,elmegreen1994a}, as indicated by observations \\citep{blitz2004a,blitz2006a,gardan2007a}. Analytical models and numerical simulations that tie star formation to the fraction of gas in the dense ISM are successful in reproducing many observational trends \\citep[e.g.,][]{elmegreen2002a,kravtsov2003a,krumholz2005a,li2005a,li2006a,tasker2006a,tasker2007a,krumholz2007a,wada2007a,tassis2007a}. Recently, several studies have explored star formation recipes based on molecular hydrogen. \\citet{pelupessy2006a} and \\citet{booth2007a} implemented models for $\\Hmol$ formation in gaseous disks and used them to study the molecular content and star formation in galaxies. However, these studies focused on the evolution of galaxies of a single mass and did not address the origin of the SK relation, its dependence on galaxy mass or structure, or its connection to trends in the local molecular fraction. Our study examines the SK relation critically, including its dependence on the structural and ISM properties of galaxies of different masses, to explain the observed deviations from the global SK relation, to investigate other connections between star formation and disk galaxy properties such as rotation or gravitational instability, and to explore how the temperature and density structure of the ISM pertains to the star formation attributes of galaxies. To these ends, we develop a model for the ISM and star formation whose key premise is that star formation on the scales of molecular clouds ($\\sim 10$~pc) is a function of molecular hydrogen density with a universal SFE per free-fall time \\citep[e.g.,][]{krumholz2007b}. Molecular hydrogen, which we assume to be a proxy for dense, star forming gas, is accounted for by calculating the local $\\Hmol$ fraction of gas as a function of density, temperature, and metallicity using the photoionization code Cloudy \\citep{ferland1998a} to incorporate $\\Hmol$-destruction by the UV radiation of local young stellar populations. We devise a numerical implementation of the star formation and ISM model, and perform hydrodynamical simulations to study the role of molecular gas in shaping global star formation relations of self-consistent galaxy models over a representative mass range. The results of our study show that many of the observed global star formation correlations and trends can be understood in terms of the dependence of molecular hydrogen abundance on the local gas volume density. We show that the physics controlling the abundance of molecular hydrogen and its destruction by the interstellar radiation field (ISRF) play a key role in shaping these correlations, in agreement with earlier calculations based on more idealized models of the ISM \\citep{elmegreen1993a,elmegreen1994a}. While our simulations focus on the connection between the molecular ISM phase and star formation on galactic scales, the formation of molecular hydrogen has also been recently studied in simulations of the ISM on smaller, subgalactic scales \\citep{glover2007a,dobbs2007a}. These simulations are complementary to the calculations presented in our study and could be used as input to improve the molecular ISM model we present. The paper is organized as follows. The simulation methodology, including our numerical models for the ISM, interstellar radiation field, and simulated galaxies, is presented in \\S \\ref{section:methodology}. The results of the simulations are presented in \\S \\ref{section:results}, where the simulated star formation relations in galactic disks and correlations of the molecular fraction with the structure of the ISM are examined. We discuss our results in \\S \\ref{section:discussion} and conclude with a summary in \\S \\ref{section:summary}. Details of our tests of numerical fragmentation in disk simulations and calculations of the model scaling between star formation, gas density, and orbital frequency are presented in the Appendices. Throughout, we work in the context of a dark-energy dominated cold dark matter cosmology with a Hubble constant $H_{0} \\approx 70\\kms\\Mpc^{-1}$. \\begin{figure*} \\figurenum{1} \\epsscale{1} \\plotone{fig1.eps} \\caption{\\label{fig:cooling} Cooling ($\\Lambda$) and heating ($\\Heating$) rates for interstellar and intergalactic gas as a function of gas density ($\\nH$), temperature ($T$), metallicity ($Z$), and interstellar radiation field (ISRF) strength ($\\Uisrf$), in units of the ISRF strength in the MW at the solar circle, as calculated by the code Cloudy \\citep{ferland1998a}. Shown are the cooling (dotted lines), heating (dashed lines) and net cooling (solid) functions over the temperature range $T=10^{2}-10^{9}~\\K$. Dense gas can efficiently cool via atomic and molecular coolants below $T=10^{4}\\K$, depending on the gas density and the strength of the ISRF. A strong ISRF can enable the destruction of $\\Hmol$ gas and thereby reduce the SFR. } \\end{figure*} ", "conclusions": "\\label{section:discussion} Results presented in the previous sections indicate that star formation prescriptions based on the local abundance of molecular hydrogen lead to interesting features of the global star formation relations. We show that the inclusion of an interstellar radiation field is critical to control the amount of diffuse $\\Hmol$ at low gas densities. For instance, without a dissociating ISRF the low mass dwarf galaxy eventually becomes almost fully molecular, in stark contrast with observations. We also show that without the dissociating effect of the ISRF our model galaxies produce a much flatter relation between molecular fraction $\\fHmol$ and pressure $\\Pext$, as can be expected from the results of \\cite{elmegreen1993a}. Including $\\Hmol$-destruction by an ISRF results in a $\\fHmol-\\Pext$ relation in excellent agreement with the observations of \\citet{wong2002a} and \\citet{blitz2004a,blitz2006a}. Our model also predicts that the relation between $\\Sigma_{\\rm SFR}$ and $\\Sigma_{\\rm gas}$ should not be universal and can be considerably steeper than the canonical value of $n_{\\rm tot}=1.4$, even if the three-dimensional Schmidt relation in molecular clouds is universal. The slope of the relation is controlled by the dependence of molecular fraction (i.e., fraction of star forming gas) on the total local gas surface density. This relation is non-trivial because the molecular fraction is controlled by pressure and ISRF strength, and can thus vary between different regions with the same total gas surface density. The relation can also be different in the regions where the disk scale-height changes rapidly (e.g., in flaring outer regions of disks), as can be seen from equation~\\ref{equation:structural_schmidt_law} \\citep[see also][]{schaye2007a}. We show that the effect of radial variations in the molecular fraction $\\fHmol$ and gas scale heights ($\\hSFR$ and $\\hgas$) on the SFR can be accounted for in terms of a structural, SK-like correlation, $\\SigmaSFR\\propto\\fHmol \\hSFR \\hgas^{-1.5} \\Sigmagas^{1.5}$, that trivially relates the local SFR to the consumption of molecular gas with an efficiency that scales with the local dynamical time. A generic testable prediction of our model is that deviations from the SK $\\SigmaSFR-\\Sigmagas$ relation are expected in galaxies or regions of galaxies where the molecular fraction is declining or much below unity. As we discussed above, star formation in the molecule-poor ($\\fHmol\\sim0.1$) galaxy M33 supports this view as its total gas SK power-law index is $\\ntot\\approx 3.3$ \\citep{heyer2004a}. While our simulations of an M33 analogue produce a steep SK power-law, the overall SFE in our model galaxies lies below that observed for M33. However, as recently emphasized by \\cite{gardan2007a}, the highest surface density regions of M33 have unusually efficient star formation compared with the normalization of the \\cite{kennicutt1998a} relation and so the discrepancy is not surprising. Given that dwarf galaxies generally have low surface densities and are poor in molecular gas, it will be interesting to examine SK relation in other small-mass galaxies. Another example of a low-molecular fraction galaxy close to home is M31, which has only a fraction $\\fHmol\\approx0.07$ of its gas in molecular form within 18 kpc of the galactic center \\citep{nieten2006a}. Our model would predict that this galaxy should deviate from the total gas SK relation found for molecular-rich galaxies. Observationally, the SK relation of M31 measured by \\citet{boissier2003a} is rather complicated and even has an increasing SFR with decreasing gas density over parts of the galaxy. Low molecular fractions can also be expected in the outskirts of normal galaxies and in the disks of low surface brightness galaxies. The latter have molecular fractions of only $\\fHmol\\lesssim 0.10$ \\citep{matthews2005a,das2006a}, and we therefore predict that they will not follow the total gas SK relation obeyed by molecule-rich, higher surface density galaxies. At the same time, LSBs do lie on the same relation between $\\Hmol$ mass and far infrared luminosity as higher surface brightness (HSB) galaxies \\citep{matthews2005a}, which suggests that the dependence of star formation on molecular gas may be the same in both types of galaxies. An alternative formulation of the global star formation relation is based on the angular frequency of disk rotation: $\\Sigma_{\\rm SFR}\\propto \\Sigmagas\\Omega$. That this relation works in real galaxies is not trivial, because star formation and dynamical time-scale depend on the local gas density, while $\\Omega$ depends on the total mass distribution {\\it within} a given radius. Although several models were proposed to explain such a correlation \\citep[see, e.g.,][for reviews]{kennicutt1998a,elmegreen2002a}, we show in \\S \\ref{section:results:sfr_rotation} and the second Appendix that the star formation correlation with $\\Omega$ can be understood as a fortuitous correlation of $\\Omega$ with gas density of $\\Omega\\propto \\rho_{\\rm gas}^{\\alpha}$, where $\\alpha\\approx 0.5$, for self-gravitating exponential disks or exponential disks embedded in realistic halo potentials. Moreover, we find that $\\SigmaSFR\\propto \\Sigmagas\\Omega$ breaks down at low values of $\\Sigmagas\\Omega$ where the molecular fraction declines, similarly to the steepening of the SK relation. Our models therefore predict that the $\\SigmaSFR\\propto\\SigmaHmol\\Omega$ relation is more robust than the $\\SigmaSFR\\propto\\Sigmagas\\Omega$ relation. An important issue related to the global star formation in galaxies is the possible existence of star formation thresholds \\citep{kennicutt1998a,martin2001a}. Such thresholds are expected to exist on theoretical grounds, because the formation of dense, star-forming gas is thought to be facilitated by either dynamical instabilities \\citep[see, e.g.,][for comprehensive reviews]{elmegreen2002a,mckee2007a} or gravithermal instabilities \\citep{schaye2004a}. We find that the two-component Toomre instability threshold that accounts for both stars and gas, $Q_{\\rm sg}<1$, works well in predicting the transition from atomic gas, inert to star formation, to the regions where molecular gas and star formation occur in our simulations. Our results are in general agreement with \\citet{li2005a,li2005b,li2006a}, who used sink-particle simulations of the dense ISM in isolated galaxies to study the relation between star formation and the development of gravitational instabilities, and with observations of star formation in the LMC \\citep{yang2007a}. Our simulations further demonstrate the importance of accounting for all mass components in the disk to predict correctly which regions galactic disks are gravitationally unstable. Given that our star formation prescription is based on molecular hydrogen, the fact that $Q_{\\rm sg}$ is a good threshold indicator may imply that gravitational instabilities strongly influence the abundance of dense, molecular gas in the disk. Conversely, the gas at radii where the disk is stable remains at low density and has a low molecular fraction. We find that in our model galaxies, the shear instability criterion of \\cite{elmegreen1993a} does not work as well as the Toomre $Q_{\\rm sg}$-based criterion. Almost all of the star formation in our model galaxies occurs at surface densities $\\Sigmagas\\gtrsim 3{\\rm\\ M_{\\odot}\\,pc^{-2}}$, which is formally consistent with the \\citet{schaye2004a} constant surface density criterion for gravithermal instability. However, as Figure~\\ref{fig:kennicutt.gas} shows, we do not see a clear indication of threshold at a particular surface density and our GD-SF models that have an effectively isothermal ISM with $T\\approx10^{4}\\K$ (and hence do not have a gravithermal instability) still show a good correlation between regions where $\\Qsg<1$ and regions where star formation operates. Our results have several interesting implications for interpretation of galaxy observations at different epochs. First, low molecular fractions in dwarf galaxies mean that only a small fraction of gas is participating in star formation at any given time. This connection between SFE and molecular hydrogen abundance may explain why dwarf galaxies are still gas rich today compared to larger mass galaxies \\citep{geha2006a}, without relying on mediation of star formation or gas blowout by supernovae. Note that a similar reasoning may also explain why large LSBs at low redshifts are gas rich but anemic in their star formation. Understanding the star formation and evolution of dwarf galaxies is critical because they serve as the building blocks of larger galaxies at high redshifts. Such small-mass galaxies are also expected to be the first objects to form large masses of stars and should therefore play an important role in enrichment of primordial gas and the cosmic star formation rate at high redshifts \\citep{hopkins2004a,hopkins2006am}. The star-forming disks at $z\\sim2$ that may be progenitors of low-redshift spiral galaxies are observed to lack centrally-concentrated bulge components \\citep{elmegreen2007a}. Given that galaxies are expected to undergo frequent mergers at $z>2$, bulges should have formed if a significant fraction of baryons are converted into stars during such mergers \\citep[e.g.,][]{gnedin2000a}. The absence of the bulge may indicate that star formation in the gas rich progenitors of these $z\\sim 2$ systems was too slow to convert a significant fraction of gas into stars. This low SFE can be understood if the high cosmic UV background, low-metallicities, and low dust content of high-$z$ gas disks keep their molecular fractions low \\citep{pelupessy2006a}, thereby inhibiting star formation over most of gas mass and keeping the progenitors of the star-forming $z\\sim 2$ disks mostly gaseous. Gas-rich progenitors may also help explain the prevalence of extended disks in low-redshift galaxies despite the violent early merger histories characteristic of $\\Lambda$CDM universes, as gas-rich mergers can help build high-angular momentum disk galaxies \\citep[][]{robertson2006c}. Mergers of mostly stellar disks, on the other hand, would form spheroidal systems. Our results may also provide insight into the interpretation of the results of \\cite{wolfe2006a}, who find that the SFR associated with neutral atomic gas in damped Lyman alpha (DLA) systems is an order of magnitude lower than predicted by the local \\citet{kennicutt1998a} relation. The DLAs in their study sample regions with column densities $N_{\\mathrm{H}}\\approx 2\\times 10^{21}\\rm\\ cm^{-2}$, or surface gas densities of $\\Sigma_{\\rm DLA}\\approx 20\\rm\\ M_{\\odot}\\,pc^{-2}$, assuming a gas disk with a thickness of $h\\approx100$~pc. Suppose the local SK relation steepens from the local relation with the slope $n_0=1.4$ to a steeper slope $n_1$ below some surface density $\\Sigma_{\\rm b}>\\Sigma_{\\rm DLA}$. Then for $\\Sigmagas<\\Sigma_{\\rm b}$, the SFR density will be lower than predicted by the local relation by a factor of $(\\Sigmagas/\\Sigma_{\\rm b})^{n_1-n_0}$. For $n_0=1.4$ and $n_1=3$ the SFR will be suppressed by a factor of $>10$ for $\\Sigmagas/\\Sigma_{\\rm b}<0.25$. Thus, the results \\cite{wolfe2006a} can be explained if the total gas SK relation at $z\\sim3$ steepens below $\\Sigmagas\\lesssim 100\\rm\\ M_{\\odot}\\, pc^{-2}$. We suggest that if the majority of the molecular hydrogen at these redshifts resides in rare, compact, and dense systems \\citep[e.g.,][]{zwaan2006a}, then both the lack of star formation and the rarity of molecular hydrogen in damped Ly$\\alpha$ absorbers may be explained simultaneously. Our results also indicate that the thermodynamics of the ISM can leave an important imprint on its density probability distribution. Each thermal phase in our model galaxies has its own log-normal density distribution. Our results thus imply that using a single lognormal PDF to build a model of global star formation in galaxies \\citep[e.g.,][]{elmegreen2002a,wada2007a} is likely an oversimplification. Instead, the global star formation relation may vary depending on the dynamical and thermodynamical properties of the ISM. We can thus expect differences in the SFE between the low-density and low-metallicity environments of dwarf and high-redshift galaxies and the higher-metallicity, denser gas of many large nearby spirals. Note that many of the results and effects we discuss above may not be reproduced with a simple 3D density threshold for star formation, as commonly implemented in galaxy formation simulations. Such a threshold can reproduce the atomic-to-molecular transition only crudely and would not include effects of the local interstellar radiation field, metallicity and dust content, etc. A clear caveat for our work is that the simulation resolution limits the densities we can model correctly. At high densities, the gas in our simulations is over-pressurized to avoid numerical Jeans instability. The equilibrium density and temperature structure of the ISM and the molecular fraction are therefore not correct in detail. Note, however, that our pressurization prescription is designed to scale with the resolution, and should converge to the ``correct'' result as the resolution improves. In any event, the simulations likely do not include all the relevant physics shaping density and temperature PDFs of the ISM in real galaxies. The results may of course depend on other microphysics of the ISM as they influence both the temperature PDF and the fraction of gas in a high-density, molecular form. Future simulations of the molecular ISM may need to account for new microphysics as they resolve scales where such processes become important. Using hydrodynamical simulations of the ISM and star formation in cosmologically motivated disk galaxies over a range of representative masses, we examine the connection between molecular hydrogen abundance and destruction, observed star formation relations, and the thermodynamical structure of the interstellar medium. Our simulations provide a variety of new insights into the mass dependence of star formation efficiency in galaxies. A summary of our methodology and results follows. \\begin{itemize} \\item[1.] A model of heating and cooling processes in the interstellar medium (ISM), including low-temperature coolants, dust heating and cooling processes, and heating by the cosmic UV background, cosmic rays, and the local interstellar radiation field (ISRF), is calculated using the photoionization code Cloudy \\citep{ferland1998a}. Calculating the molecular fraction of the ISM enables us to implement a prescription for the star formation rate (SFR) that ties the SFR directly to the molecular gas density. The ISM and star formation model is implemented in the SPH/N-body code GADGET2 \\citep{springel2005c} and used to simulate the evolution of isolated disk galaxies. \\item[2.] We study the correlations between gas surface density ($\\Sigmagas$), molecular gas surface density ($\\SigmaHmol$), and SFR surface density ($\\SigmaSFR$). We find that in our most realistic model that includes heating and destruction of $\\Hmol$ by the interstellar radiation field, the power law index of the SK relation, $\\SigmaSFR\\propto\\Sigmagas^{\\ntot}$, (measured in annuli) varies from $\\ntot\\sim2$ in massive galaxies to $\\ntot\\gtrsim4$ in small mass dwarfs. The corresponding slope of the $\\SigmaSFR\\propto\\SigmaHmol^{\\nmol}$ molecular-gas Schmidt-Kennicutt relation is approximately the same for all galaxies, with $\\nmol\\approx 1.3$. These results are consistent with observations of star formation in different galaxies \\citep[e.g.,][]{kennicutt1998a,wong2002a,boissier2003a,heyer2004a,boissier2007a,kennicutt2007a}. \\item[3.] In our models, the SFR density scales as $\\SigmaSFR \\propto \\fHmol h_{\\SFR} h_{\\gas}^{-1.5} \\Sigmagas^{1.5}$, where $h_{\\gas}$ is the scale-height of the ISM and $h_{\\SFR}$ is the scale-height of star-forming gas. The different $\\SigmaSFR-\\Sigmagas$ relations in galaxies of different mass and in regions of different surface density in our models therefore owe to the dependence of molecular fraction $\\fHmol$ and scale height of gas on the gas surface density. \\item[4.] We show that the $\\SigmaSFR\\propto\\Sigmagas\\Omega$ and $\\SigmaSFR\\propto\\SigmaHmol\\Omega$ correlations describe the simulations results well where the molecular gas and total gas densities are comparable, while the simulations deviate from $\\SigmaSFR\\propto\\Sigmagas\\Omega$ \\citep[e.g.,][]{kennicutt1998a} at low $\\Sigmag$ owing to a declining molecular fraction. We demonstrate that these relations may owe to the fact that the angular frequency and the disk-plane gas density are generally related as $\\Omega\\propto\\sqrt{\\rho}$ for exponential disks if the potential is dominated by either the disk, a \\cite{navarro1996a} halo, \\cite{hernquist1990a} halo, or an isothermal sphere. The correlation of $\\SigmaSFR$ with $\\Omega$ is thus a secondary correlation in the sense that $\\Omega\\propto\\sqrt{\\rho}$ is set during galaxy formation and $\\Omega$ does not directly influence star formation. \\item[5.] The role of critical surface densities for shear instabilities ($\\SigmaA$) and \\cite{toomre1964a} instabilities ($\\SigmaQ$) in star formation \\citep[e.g.,][]{martin2001a} is examined in the context of the presented simulations. We find that the two-component Toomre instability criterion $\\Qsg<1$ is an accurate indicator of the star-forming regions of disks, and that gravitational instability and star formation are closely related in our simulations. Further, the $\\Qsg$ criterion works even in simulations in which cooling is restricted to $T>10^4$~K where gravithermal instability cannot operate. \\item[6.] Our simulations that include $\\Hmol$-destruction by an ISRF naturally reproduce the observed scaling $\\fHmol\\propto\\Pext^{0.9}$ between molecular fraction and external pressure \\citep[e.g.,][]{wong2002a,blitz2004a,blitz2006a}, but we find that simulations without an ISRF have a weaker scaling $\\fHmol\\propto \\Pext^{0.4}$. We calculate how the connection between the scalings of the gas surface density, the stellar surface density, and the ISRF strength influence the $\\fHmol-\\Pext$ relation in the ISM, and show how the simulated scalings reproduce the $\\fHmol-\\Pext$ relation even as the power-law index of the total gas Schmidt-Kennicutt relation varies dramatically from galaxy to galaxy. \\item[7.] We present a method for mitigating numerical Jeans fragmentation in Smoothed Particle Hydrodynamics simulations that uses a density-dependent pressurization of gas on small scales to ensure that the Jeans mass is properly resolved, similar to techniques used in grid-based simulations \\citep[e.g.,][]{truelove1997a,machacek2001a}. The gas internal energy $u$ at the Jeans scale is scaled as $u\\propto\\mJeans^{-2/3}$, where $\\mJeans$ is the local Jeans mass, to ensure the Jeans mass is resolved by some $\\NJeans$ number of SPH kernel masses $2\\Nneigh m_{\\mathrm{SPH}}$, where $\\Nneigh$ is the number of SPH neighbor particles and $m_{\\mathrm{SPH}}$ is the gas particle mass. For the simulations presented here, we find the \\cite{bate1997a} criterion of $\\NJeans=1$ to be insufficient to avoid numerical fragmentation and that $\\NJeans\\sim15$ provides sufficient stability against numerical fragmentation over the time evolution of our simulations. Other simulations may have more stringent resolution requirements \\citep[e.g.,][]{commercon2008a}. We also demonstrate that isothermal galactic disks with temperatures of $T=10^{4}\\K$ may be susceptible to numerical Jeans instabilities at resolutions common in cosmological simulations of disk galaxy formation, and connect this numerical effect to possible angular momentum deficiencies in cosmologically simulated disk galaxies. \\end{itemize} The results of our study indicate that star formation may deviate significantly from the relations commonly assumed in models of galaxy formation in some regimes and that these deviations can be important for the overall galaxy evolution. Our findings provide strong motivation for exploring the consequences of such deviations and for developing further improvements in the treatment of star formation in galaxy formation simulations.\\\\[2mm]" }, "0710/0710.2428_arXiv.txt": { "abstract": "{ A recent analysis of cosmic-ray data from a space borne experiment by the AMS collaboration supports the observation of an excess in the cosmic-ray positron spectrum by previous balloon experiments. The combination of the various experimental data establishes a deviation from the expected background with a significance of more than four standard deviations. The observed change in the spectral index cannot be explained without introducing a new source of positrons. When interpreted within the MSSM a consistent description of the antiproton spectrum, the diffuse gamma-ray flux and the positron fraction is obtained which is compatible with all other experimental data, including recent WMAP data. \\PACS{ {98.70.Sa}{Cosmic rays} \\and {95.35.+d}{Dark Matter} \\and {11.30.Pb}{Supersymmetry} } % } % ", "introduction": "Among the cosmic-ray species, antiparticles and diffuse $\\gamma$-rays are of particular interest because they are produced secondarily in hadronic interactions of protons and nuclei with the interstellar medium at low rates. Their small abundance makes them a sensitive probe for the existence of additional -- and possibly exotic -- cosmic-ray sources which would be visible as an excess of particles above conventional expectations. One of the most important unsolved questions in modern cosmology is the nature of dark matter. The most promising dark matter candidate is the weakly interacting lightest neutralino, $\\chi_1^0$, predicted by supersymmetric extensions to the standard model of particle physics. The annihilation of neutralinos might constitute an additional primary source of particles with a unique spectral shape which would be determined by the parameters of supersymmetry, allowing to put constraints on new physics beyond the standard model. A recent reanalysis of the data from the \\AMS{} spectrometer~\\cite{aguilar07a} supports the observation of an excess of cosmic-ray positrons by the HEAT experiments~\\cite{beatty04a}. In this work, we discuss the combined results on the cosmic-ray positron fraction $e^+ / (e^+ + e^-)$. Assuming that dark matter is largely constituted by neutralinos, we determine the cosmic-ray preferred parameter space of the minimal supersymmetric standard model (MSSM) from a simultaneous fit to the cosmic-ray positron, antiproton and diffuse $\\gamma$-ray data. ", "conclusions": "\\label{sec:conclusions} In this work, the combined recent experimental results on the cosmic-ray positron fraction have been presented. The data exhibit an excess of positrons above energies of 6\\+\\GeV{} which cannot be explained by purely secondary positron production alone and thus requires an additional primary source of positrons. In this work, we interpret this source to be the annihilation of supersymmetric neutralinos constituting dark matter. A simultaneous fit to the cosmic-ray positron, antiproton and $\\gamma$-ray data shows that, for particular sets of the MSSM parameters, this hypothesis gives a fully consistent description of the cosmic-ray spectra which is compatible with all other experimental data. We find that the cosmic-ray data clearly prefer the focus point region of the MSSM parameter space but reveal almost no sensitivity to $\\tan\\beta$." }, "0710/0710.5452_arXiv.txt": { "abstract": "We use a series of ray-tracing experiments to determine the magnification distribution of high-redshift sources by gravitational lensing. We determine empirically the relation between magnification and redshift, for various cosmological models. We then use this relation to estimate the effect of lensing on the determination of the cosmological parameters from observations of high-$z$ supernovae. We found that, for supernovae at redshifts $z<1.8$, the effect of lensing is negligible compared to the intrinsic uncertainty in the measurements. Using mock data in the range $1.82$. Therefore if supernovae up to these redshifts were ever discovered, it is still the ones in the range $0.31$ would result in a increase in $F$, and an underestimation of $d_L$. Estimating the effect of lensing on the statistics of high-$z$ supernovae is a complex problem. Using either an analytical model or ray-tracing simulations, we can estimate the effect of lensing of a large number of sources in a statistical sense. We would then need to redo the error analysis on the SNe data to include in a consistent way the effect of lensing. This would be a very complex task, and in this paper we have chosen a much simpler approach. {\\it Our goal is not to obtain a precise estimate of the error introduced by lensing, but rather to assess the importance of this effect: is it dominant, important, or negligible, and for what range of redshift? and how does it affect the discrimination between different cosmological models?\\/} To answer these questions, we take at face value the published results of Type~Ia SNe, including their error bars which account for every source of uncertainty but gravitational lensing. Then, we include {\\it a posteriori\\/} the effect of lensing, to estimate the change in the errors. This approach is not rigorous at all, and does not constitute a substitute for a rigorous treatment of the errors. But it has the great advantage of simplicity. We do not have to redo the detailed error analysis performed by the high-redshift SNe groups, and, more importantly, our conclusions will not be tied to any particular sample or particular data reduction and error analysis technique used by any particular group. We are seeking to make generic statements about the importance of lensing (or lack of) that are relevant to any current or future sample of high-$z$ SNe. The lensing of distant supernovae has been the focus of several recent studies. In an early study, \\citet{wambsganssetal97} used ray-tracing experiments to estimate the effect of weak lensing on the determination of the deceleration parameter $q_0$. \\citet{md05}, \\citet{dv05}, and \\citet{mv06} focused on SNe as a mean to study the nature of weak lensing. The issue of determining the cosmological parameters for distant SNe, and how this determination is affected by lensing, was addressed by \\citet{wang05} who used semi-analytical models to determine the magnification distribution function, \\citet{hl05} who used Monte Carlo ray-tracing simulations to study the effect of weak and strong lensing, and \\citet{gunnarssonetal06} and \\citet{jonssonetal06}, who estimated the effect on lensing along individual lines of sight by considering the properties of foreground galaxies in the same direction. These various studies concluded that the effect of lensing on current determinations of the cosmological parameters is small. \\citet{alderingetal06} discussed the effect of gravitational lensing on a population of SNe at $z>1.7$. What distinguishes our approach is mostly its simplicity. Our calculations depend on very few assumptions, and this implies a certain amount of robustness to our results. Even though we rely on numerical simulations, this work should be regarded as a back-of-the-envelope calculation, whose purpose is to obtain a qualitative estimate of the effect of lensing on the determination of cosmological parameters by distant SNe. Using ray-tracing experiments, rather than a semi-analytical approach, enables us to extend our study to redshifts much higher than the ones considered by \\citet{wang05} and \\citet{hl05}. This paper is organized as follow: In \\S2, we describe our calculation of the magnification distribution $P(\\mu)$, and how to estimate that distribution at any redshift $z$. In \\S3, we describe the real and mock samples of supernovae we use for our calculations. Results are presented in \\S4. In \\S5, we address various observational issues. Summary and conclusion are presented in \\S6. ", "conclusions": "We have performed a series of ray-tracing experiments using a multiple lens-plane algorithm. We have determined the distributions of magnifications $P(\\mu)$ for sources in the redshift range $02$ could be quite significant, and must be understood before such SNe could be used to constrain cosmological models. Furthermore, the open CDM and $\\Lambda$CDM are difficult to distinguish at that redshift. We showed that, even if SNe at redshift $z\\sim8$ were ever discovered, it is the SNe in the range $z=0.3-1$ that would still provide the best discriminant between these two models. The data at that redshift already exist, and they support the $\\Lambda$CDM model." }, "0710/0710.3369_arXiv.txt": { "abstract": " ", "introduction": "It is thought that the decay of the magnetic field in \\emph{magnetars} is the main source of its X-ray and $\\gamma$-ray luminosity since these objects appear to radiate substantially more power than available from the energy rotational loss \\cite{TD-96}. The spontaneous decay of the magnetic field could occur through \\emph{ambipolar diffusion}, \\emph{Hall drift} and \\emph{ohmic diffusion}, which are non-ideal magnetohydrodynamics processes that occur in thousands of years, compared to dynamical sound and Alfv\\'en time scales of milliseconds to seconds. Ambipolar diffusion promote a dissipative magnetic field advection, through the movement of a bulk of charged particles relative to the neutrons, the \\emph{Hall drift} is a non-dissipative advection of the magnetic field caused by the electrical current associated with it, and the magnetic field ohmic diffusion is a dissipative process caused by the electrical resistivity. The time scales of these process were estimated in Ref.~2. For classical pulsar magnetic field strengths, these were found to be longer than the lifetimes of these stars, and therefore unlikely to be observationally important. For the case of magnetars, it was found that magnetic field decay by these processes might be occurring \\cite{TD-96,ACT-04}. However, a full understanding of these processes, their interactions, and their effectiveness in neutron stars is still lacking. The work of \\cite{GR-92} was analytical, and therefore useful to identify general processes and identify the relevant time scales, but not to address the action of the identified processes in their full nonlinear development, and their interaction with each other. The full evolution of the magnetic field can only be addressed by numerical simulations. Recent ideal three-dimensional single-fluid magnetohydrodynamics simulations showed that, in a stably stratified star, a complicated, random, initial field generally evolves on a short, Alfv\\'en-like time scale to a relatively simple, roughly axisymmetric, large-scale configuration containing a toroidal and a poloidal component of comparable strength, both of which are required in order to stabilize each other \\cite{BS-04}. It is interesting to study the effects of the non-ideal processes studied in Ref.~2 on the evolution of this configuration. As a first step in this direction, here we simulate the decay of a magnetic field, including the effects studied in Ref.~2, in a system where the magnetic field points in one Cartesian direction but varies only along an orthogonal direction. It is shown that the magnetic field evolves through different quasi-equilibrium states and we estimate the characteristic time scales where these quasi-equilibria occur. ", "conclusions": "Using numerical simulations in one dimension, we studied the effects of some non-ideal MHD processes on the magnetic field evolution in neutron stars. We found that the system evolves through succesive quasi-equilibria, and we estimated the characteristic time scales on which these quasi-equilibria occur. \\begin{center} Acknowledgments \\end{center} We acknowledge financial support through FONDECYT postdoctoral project 3060103, Gemini project 32070014, and regular FONDECYT projects 1060644 and 1070854. We also thank the ESO-Chile Mixed Committee." }, "0710/0710.4549_arXiv.txt": { "abstract": "We investigate the chemical abundances of NGC\\,3603 in the Milky Way, of 30\\,Doradus in the Large Magellanic Cloud, and of N\\,66 in the Small Magellanic Cloud. Mid-infrared observations with the Infrared Spectrograph onboard the Spitzer Space Telescope allow us to probe the properties of distinct physical regions within each object: the central ionizing cluster, the surrounding ionized gas, photodissociation regions, and buried stellar clusters. We detect [S\\3], [S\\4], [Ar\\3], [Ne\\2], [Ne\\3], [Fe\\2], and [Fe\\3] lines and derive the ionic abundances. Based on the ionic abundance ratio (Ne\\3/H)/(S\\3/H), we find that the gas observed in the MIR is characterized by a higher degree of ionization than the gas observed in the optical spectra. We compute the elemental abundances of Ne, S, Ar, and Fe. We find that the $\\alpha$-elements Ne, S, and Ar scale with each other. Our determinations agree well with the abundances derived from the optical. The Ne/S ratio is higher than the solar value in the three giant H\\2\\ regions and points toward a moderate depletion of sulfur on dust grains. We find that the neon and sulfur abundances display a remarkably small dispersion (0.11\\,dex in 15 positions in 30\\,Doradus), suggesting a relatively homogeneous ISM, even though small-scale mixing cannot be ruled out. ", "introduction": "\\label{sec:intro} Giant H\\2\\ regions are ideal laboratories to understand the feedback of star-formation on the dynamics and energetics of the interstellar medium (ISM). Supernov{\\ae} and stellar winds arising in such regions are reponsible for producing shocks, destroying dust grains and molecules, while compressing molecular clouds and triggering subsequent star-formation. They also allow the release of newly synthetized elements into the ISM, altering its metallicity. In order to study the star-formation properties as a function of the environment, we observed three giant H\\2\\ regions spanning a wide range of physical conditions (gas density, mass, age) and chemical properties (metallicity) with the Spitzer Space Telescope (Werner et al.\\ 2004). Observations are part of the GTO program PID\\#63. The regions are NGC\\,3603 in the Milky Way, 30\\,Doradus (hereafter 30\\,Dor) in the Large Magellanic Cloud (LMC), and N\\,66 in the Small Magellanic Cloud (SMC). The scope of this program is to address crucial issues such as the destruction of complex molecules by energetic photons arising from massive stars, the polycyclic aromatic hydrocarbon (PAH) abundance dependence on metallicity, or conditions that lead to the formation/disruption of massive stellar clusters. Photometry with Spitzer/IRAC (Fazio et al.\\ 2004) has been performed and will be discussed in Brandl et al.\\ (in preparation). The brightest mid-infrared (MIR) regions (knots, stellar clusters, shockfronts, ...) were followed spectroscopically with the Infrared Spectrograph (IRS; Houck et al.\\ 2004). In Lebouteiller et al.\\ (2007), we analyzed the spatial variations of the PAH and fine-structure line emission across individual photodissociation regions (PDRs) in NGC\\,3603. The two other regions will be investigated the same way in follow-up papers (Bernard-Salas et al.\\ in preparation; Whelan et al.\\ in preparation). In this paper, we introduce the full IRS dataset (low- and high-resolution) of the giant H\\2\\ regions and we derive their chemical abundances. A subsequent paper will be focused on the study of molecules and dust properties (Lebouteiller et al.\\ in preparation). Elemental abundances in H\\2\\ regions are historically derived from optical emission-lines. Large optical telescopes, together with sensitive detectors makes it possible to determine the chemical composition of very faint H\\2\\ regions. Because of dust extinction, optical spectra only observe ionized gas toward sighlines with low dust content. In this view, the MIR range allows analyzing denser lines of sight, with possibly different chemical properties because of small-scale mixing and/or differential depletion on dust grains. MIR emission-lines constitute the only way to measure abundances in more obscure regions, and these abundances ought to be compared to abundances from the optical range. Although the optical domain gives access to some of the most important elements to constrain nucleosynthethic and stellar yields (C, N, O, Ne, S, Ar, Fe), it does not include some essential ionization stages necessary for abundance determinations of certain elements, such as S\\4\\ or Ne\\2. The MIR range enables the abundance determination of Ne, S, and Ar, with the most important ionization stages observed. Iron abundance can be also determined from MIR forbidden emission-lines, but with considerably larger uncertainty due to ionization corrections. Finally, it must be stressed that abundance determinations in the optical are more sensitive to the electronic temperature ($T_e$) determination as compared to the MIR range. The effect of $T_e$ on abundances determinations is a significant source of error in optical abundance results. Wu et al.\\ (2007) recently studied a sample of blue compact dwarf galaxies (BCDs) with the IRS and found a global agreement between abundances derived from the optical and those derived from the MIR. This suggests that the dense lines of sight probed in the MIR have a similar chemical composition as unextincted lines of sight and/or dense regions with possibly peculiar abundances do not contribute significantly to the integrated MIR emission-line spectrum. MIR abundances of the BCDs were calculated using mostly H$\\beta$ or H$\\alpha$ lines from the optical as tracer of the hydrogen content, with significant uncertainties from aperture corrections, or different observed regions because of extinction. The present sample of giant H\\2\\ regions provides the unique opportunity to measure accurate abundances, with a signal-to-noise ratio sufficiently high to observe directly the H\\1\\ recombination line at 12.37\\mic. We provide abundances of Ne, S, Ar, and Fe toward lines of sight with different physical properties (PDRs, ionized gas, embedded source, stellar cluster, ...) within each giant H\\2\\ region. We first present the sample of the three giant H\\2\\ regions in \\S\\ref{sec:pres}. The data reduction and analysis are discussed in \\S\\ref{sec:observations}. We infer the ion abundances in \\S\\ref{sec:ionicab}. Elemental abundances are determined in \\S\\ref{sec:eleab} and are discussed in \\S\\ref{sec:discussion}. ", "conclusions": "We analyzed the chemical abundances in the ISM of three giant H\\2\\ regions, NGC\\,3603 in the Milky Way, 30\\,Dor in the LMC, and N\\,66 in the SMC using the MIR lines observed with the IRS onboard Spitzer. \\begin{itemize} \\item Our observations probe the ISM toward various physical regions, such as stellar clusters, ionized gas, photodissociation regions, and deeply embedded MIR bright sources. The spectra show the main ionization stages of neon, sulfur, and argon in the ionized gas. We also detect [Fe\\2] and [Fe\\3] lines. \\item Ionic abundances of Ne\\2, Ne\\3, S\\3, S\\4, Ar\\2, Ar\\3, Fe\\2, and Fe\\3\\ were derived. The internal variation of electron density across a region has no impact on the ionic abundance determination. On the other hand, we find that electron temperature uncertainties and/or intrinsic variations could be responsible for an error of 20\\% at most on the abundance determinations. Based on the (Ne\\3/H)/(S\\3/H) ionic abundance ratio, we find that the optical spectra probe a gas with a degree of ionization equal to or higher than the gas probed in the MIR. \\item Elemental abundances were determined from the ionic abundances. No ionization corrections were needed, except for iron. We find that neon, sulfur, and argon scale with each other, which is expected from stellar yields. Abundances do not show any dependence on the physical region (PDR, stellar cluster, embedded region, ...). \\item The Ne/S ratio is larger than the solar value, and suggests that sulfur could be depleted onto dust grains. The sulfur abundance in the MIR agrees best with the lowest optical determinations, which is likely due to uncertainties in the abundance determinations. \\item Iron abundance shows a larger uncertainty than Ne/H, S/H, and Ar/H. The comparison of iron and neon abundances hints at significant depletion of iron onto dust grains at large metallicities. The agreement with the optical determination of Fe/H indicates however that there is no differential depletion on dust grains between the gas probed in the MIR and in the optical. \\item Fe/H is found to be spectacularly large in one position, corresponding to a supernova remnant. This strongly suggest that iron atoms have been released from dust grains due to schocks from the SN. \\item The metallicity of NGC\\,3603 agrees with the Galactic abundance gradient. The metallicities of 30\\,Doradus and N\\,66 agree well with those of the PNe in their respective host galaxies. These findings suggest that the giant H\\2\\ regions did not experience a significant metal enrichment for at least 1\\,Gyr. If enrichement occured, the metallicity was altered by less than a factor of two. \\item Neon and sulfur abundances show remarkably little dispersion in the three H\\2\\ regions (e.g., 0.11\\,dex dispersion in 15 positions in 30\\,Dor). Small-scale mixing is apparently effective, abundance fluctuations are smaller than $\\sim$55\\%. However, internal variations of the abundances are likely to be on the order of $\\lesssim$5\\%, and determining their existence would require a significant improvement of the data quality and of the method to be evidenced. \\end{itemize}" }, "0710/0710.1193_arXiv.txt": { "abstract": "\\small Small perturbations in spherical and thin disk stellar clusters surrounding massive a black hole are studied. Due to the black hole, stars with sufficiently low angular momentum escape from the system through the loss cone. We show that stability properties of spherical clusters crucially depend on whether the distribution of stars is monotonic or non-monotonic in angular momentum. It turns out that only non-monotonic distributions can be unstable. At the same time the instability in disk clusters is possible for both types of distributions. ", "introduction": "The study of the gravitational loss-cone instability, a far analog of the plasma cone instability, has begun with the work of V. Polyachenko (1991), in which a simplest analytical model of thin disk stellar cluster has been treated. The interest to the problem of stability of stellar clusters has been revived recently by detailed investigation by Tremaine (2005) and Polyachenko, Polyachenko, Shukhman, (2007; henceforth, Paper I) of low mass clusters around massive black holes. The both papers have considered stability of small amplitude perturbations of stellar clusters of disk-like and spherical geometry. Tremaine (2005) has shown using Goodman's (1988) criterion that thin disks with symmetric DFs over angular momentum and empty loss cone are generally unstable. By contrast, analyzing perturbations with spherical numbers $l=1$ and $l=2$, he deduced that spherical clusters with monotonically increasing DF of angular momentum should be generally stable. Later we demonstrated (see Paper I) that spherical systems with non-monotonic distributions may be unstable for sufficiently small-scale perturbations $l \\ge 3 $, while the harmonics $l=1,2$ are always stable. For the sake of convenience, we have used two assumptions. The first one is that the Keplerian potential of the massive black hole dominates over a self-gravitating potential of the stellar cluster (which does not mean that one can neglect the latter). Then the characteristic time of system evolution is of the order of the orbit precessing time, which is slow, compared to typical dynamical (free fall) time. Since a star makes many revolutions in its almost unaltered orbit, we can regard it as to be ``smeared out'' along the orbit in accordance with passing time, and study evolution of systems made of these extended objects. The second assumption is a so called {\\it spoke approximation}, in which a system consists of near-radial orbits only. This approximation was earlier suggested by one of the authors (Polyachenko 1989, 1991). The spoke approximation reduces the problem to a study of rather simple analytical characteristic equations controlling small perturbations of stellar clusters. There are two questions that naturally arise in this context. First: Does the instability remain when abandoning the assumption of strong radial elongation of orbits? Second: Does the instability occur in spheres with monotonically increasing distributions in angular momentum if one consider smaller-scale perturbations with $l \\ge 3$? The aim of the paper is to provide answers to these questions. To achieve the task we use semi-analytical approach based on analysis of integral equations for slow modes elaborated recently in Polyachenko (2004, 2005) for thin disks, and in Paper I for spherical geometry. Following Paper I, we shall restrict ourselves to studying monoenergetic models with DFs in the form \\begin{align}\\label{eq:1.1} F(E,L)=A\\,\\delta(E-E_0)\\,f(L). \\end{align} The models specified by function $f(L)$ are suitable for studying the effects of angular momentum distribution on gravitational loss-cone instability. On the other hand, the Dirac $\\delta$-function permits one to reduce the integral equations for slow modes to one-dimensional integral equations, and to advance substantially in analytical calculations. Several arguments can be brought in favour of our simplified approach. First of all, the Lynden-Bell derivative (see Paper I, eq. 4.7) of the DF with respect to angular momentum $L$, keeping $J = L + I_1$ constant (here $I_1$ is the radial action) in the limit where the slow mode approximation is applicable, can be replaced by a derivative, keeping energy $E$ constant: $$ \\left(\\frac{\\p F}{\\p L} \\right)_{LB} = \\Omega_\\textrm{pr} \\left(\\frac{\\p F}{\\p E} \\right)_L + \\left(\\frac{\\p F}{\\p L} \\right)_E \\approx \\left(\\frac{\\p F}{\\p L} \\right)_{E}, $$ because $\\Omega_\\textrm{pr}$ is small. Thus, the derivative over energy is not included into the slow integral equation, and one can loosely say, that dependence on energy is only parametric. Another argument is that the results of independent study by Tremaine (2005), who used a non-monoenergetic DF, are in agreement with our conclusions. Section 2 is devoted to spheres, Section 3 -- to thin disks with symmetric DFs. The sections are organized alike. In the beginning we derive integral equations for initial distribution functions in the form (\\ref{eq:1.1}). Then follow analytical and numerical investigations of these equations. We demonstrate that by contrast to the case of near-Keplerian sphere, the loss-cone instability in disks takes place even for the monotonic DF, $df/d|L|>0$, provided the precession is retrograde and the loss cone is empty: $f(0)=0$. Sec. 2 is complimented by stability analysis of models with circular orbits, which of course doesn't belong to the class of monoenergetic models of (\\ref{eq:1.1}) type. In the last, Section 4, we discuss the results and some perspectives of further studies. ", "conclusions": "We have studied the stability of the spherically-symmetric and thin disk stellar clusters around a massive black hole. We conclude that stability properties of spherical clusters depend crucially on monotonity of initial distribution functions, while thin disk clusters are almost always unstable. If the initial distribution of the spherical cluster is monotonic, the cluster is most likely to be stable. This conclusion was first made in Tremaine (2005), where stability of $l=1$ mode was generally proved, and $l=2$ was tested numerically. We confirm this conclusion by considering a number of monotonic distributions for modes with arbitrary $l$. Besides, we have checked distributions obtained from monotonic ones by making them vanish quickly but smoothly at circular orbits. These models were also stable. However, a general proof of stability for any monotonic distributions was not yet found. Spherical clusters with the non-monotonic DFs should be generally affected by the gravitational loss-cone instability. The instability was first found in our Paper I using a simplification of systems with near-radial orbits. In the Sec. 2 we show that this instability is due to just non-monotony of distributions over angular momentum, the orbits may not necessary be near-radial. In our opinion, both monotonic and non-monotonic distributions are important for possible applications to real stellar clusters around black holes. The DFs monotonically increasing from the loss cone radius up to circular orbits are formed naturally due to two-body collisions of stars. It follows from numerical experiments (see, e.g., Cohn and Kulsrud, 1978), which predict establishment of such distributions after a characteristic time for collisional relaxation. These distributions may be approximated by the formula $F\\propto \\ln \\bigl(L/L_{\\rm min}\\bigr)$. Such a slowly increasing function is, in fact, predetermined by the boundary conditions imposed in the cited numerical study and some other investigations. Indeed, the vanishing condition at $L=L_{\\rm min}$, and the matching condition to isotropic (Maxwellian) distribution, $F=F(E)$, at the boundary $E=E_{\\rm bound}=0$ of the phase space $(E,L)$ (boundary separates stars which is gravitationally coupled to the black hole from the others) is required. The last condition means the asymptotic (when $E \\to E_{\\rm bound}$) independence of the function $F(E,L)$ on the momentum $L$. So monotonic, or logarithmic, dependence of type of (\\ref{eq:3.2}) is quite reasonable. The non-monotonic distributions are also real. If the cluster, is formed, for example, as a result of the collisionless collapse (several free fall times), then it remains collisionless for a long timescale of collisional relaxation (see, e.g., Merritt \\& Wang, 2005). In principle, the system can have almost arbitrary DF both in the energy and in the angular momentum. During the collapse, a typical non-monotonic distribution of stars over the angular momentum, with empty loss cone and maximum at some value $L=L_{\\ast}$, is formed. In Paper I we argued that stability properties of such a distribution is effectively analogous to one of typical plasma distributions of the ``beam-like'' type. But they can readily become unstable, as it is well-known in plasma physics (and also confirmed by direct stability study of corresponding stellar systems in Paper I). It is possible (as it is often so in plasma) that for the time of collisionless behavior, DF can undergo a dramatic change from its initial form. In particular, the collective flux of stars into the loss cone caused by the instability could, in principle, lead to the formation of a considerable part of the black hole. Checking of such possibilities is the most urgent task for future studies of unstable {\\it non-monotonic} models. Since spherically-symmetric models with the {\\it monotonic} DF are apparently stable, but analogous disk systems are unstable (see Tremaine 2005 and Sec. 3), a critical flatness of ellipsoid models at which the instability begins is expected. Study of such systems, as well as systems with more complex triaxial ellipsoids can be performed using numerical simulations." }, "0710/0710.1700_arXiv.txt": { "abstract": "The general world model for homogeneous and isotropic universe has been proposed. For this purpose, we introduce a global and fiducial system of reference (world reference frame) constructed on a $5$-dimensional space-time that is embedding the universe, and define the line element as the separation between two neighboring events that are distinct in space and time, as viewed in the world reference frame. The effect of cosmic expansion on the measurement of physical distance has been correctly included in the new metric, which differs from the Friedmann-Robertson-Walker metric where the spatial separation is measured for events on the hypersurface at a constant time while the temporal separation is measured for events at different time epochs. The Einstein's field equations with the new metric imply that closed, flat, and open universes are filled with positive, zero, and negative energy, respectively. The curvature of the universe is determined by the sign of mean energy density. We have demonstrated that the flat universe is empty and stationary, equivalent to the Minkowski space-time, and that the universe with positive energy density is always spatially closed and finite. In the closed universe, the proper time of a comoving observer does not elapse uniformly as judged in the world reference frame, in which both cosmic expansion and time-varying light speeds cannot exceed the limiting speed of the special relativity. We have also reconstructed cosmic evolution histories of the closed world models that are consistent with recent astronomical observations, and derived useful formulas such as energy-momentum relation of particles, redshift, total energy in the universe, cosmic distance and time scales, and so forth. It has also been shown that the inflation with positive acceleration at the earliest epoch is improbable. ", "introduction": "\\label{sec:intro} The main goal of modern cosmology is to build a cosmological model that is consistent with astronomical observations. To achieve this goal, tremendous efforts have been made both on theories and on observations since the general theory of relativity was developed. So far the most successful model of the universe is the Friedmann-Robertson-Walker (FRW) world model \\cite{fri22,fri24,rob29,wal35}. The FRW world model predicts reasonably well the current observations of the cosmic microwave background (CMB) radiation and the large-scale structures in the universe. The precisely determined cosmological parameters of the FRW world model imply that our universe is consistent with the spatially flat world model dominated by dark energy and cold dark matter ($\\Lambda\\textrm{CDM}$) with adiabatic initial condition driven by inflation \\cite{spergel07,tegmark06}. Although the flat FRW world model is currently the most reliable physical world model, one may have the following fundamental questions on the nature of the FRW world model. First, mathematically, if a space-time manifold is flat, then the Riemann curvature tensor should vanish, and vice versa. However, the Riemann curvature tensor of the flat FRW world model does not vanish unless the cosmic expansion speed and acceleration are zeros, which implies that the physical space-time of the flat FRW world is not geometrically flat but curved. Only its spatial section at a constant time is flat. Secondly, the cosmic evolution equations of the FRW world model can be derived from an application of the Newton's gravitation and the local energy conservation laws to the dynamical motion of an expanding sphere with finite mass density \\cite{milne34,mccrea34}. Besides, the Newton's gravitation theory has been widely used to mimic the non-linear clustering of large-scale structures in the universe even on the horizon-sized $N$-body simulations \\cite{colberg2000,park05}. On large scales, the close connection between the FRW world model and the Newton's gravitation law is usually attributed to the fact that the linear evolution of large-scale density perturbations satisfies the weak gravitational field condition. Recently, Hwang and Noh \\cite{hwang06} show that the relativistic fluid equations perturbed to second order in a flat FRW background world coincide exactly with the Newtonian results, and prove that the Newtonian numerical simulation is valid in all cosmological scales up to the second order. However, one may have a different point of view that the Newton's gravitational action at a distance appears to be valid even on the super-horizon scales in the FRW world just because the world model does not reflect the full nature of the relativistic theory of gravitation. Thirdly, according to the FRW world model, the universe at sufficiently early epoch ($z \\gtrsim 1000$) is usually regarded as flat since the curvature parameter contributes negligibly to the total density. The present non-flat universe should have had the density parameter approaching to $\\Omega = 1$ with infinitely high precision just after the big-bang (flatness problem). On the other hand, if we imagine the surface of an expanding balloon with positive curvature, then the curvature of the surface is always positive and becomes even higher as the balloon is traced back to the earlier epoch when it was smaller. This prediction from the common sense contradicts the FRW world model. Observationally, the flat $\\Lambda\\textrm{CDM}$ universe is favored by the recent joint cosmological parameter estimation using the Wilkinson Microwave Anisotropy (WMAP) CMB \\cite{hinshaw07,page07}, large-scale structures \\cite{cole05,tegmark04}, type Ia supernovae (SNIa; \\cite{riess07,wood07}), Hubble constant \\cite{freedman01,macri06,sandage06}, baryonic oscillation data \\cite{eisen05}, and so on. However, the WMAP CMB data alone is more compatible with the non-flat FRW world model ($\\S7.3$ of \\cite{spergel07} and Table III of \\cite{tegmark06}). Besides, some parameter estimations using SNIa data or angular size-redshift data of distant radio sources alone suggest a possibility of the closed universe \\cite{clocchi06,jackson06}. The combinations of the WMAP plus the SNIa data or the Hubble constant data also imply the possibility of the closed universe, giving curvature parameters $\\Omega_k = -0.011\\pm 0.012$ and $\\Omega_k = -0.014\\pm 0.017$, respectively \\cite{spergel07}, although the estimated values are still consistent with the flat FRW world model. The questions above and the observational constraints on the cosmological model may bring about possibilities of non-flat or non-FRW world models. Interestingly, Einstein claimed that our universe is spatially bounded or closed \\cite{ein22}. The primary reason for his preference to the closed universe is because Mach's idea \\cite{mach93,misner73} that the inertia depends upon the mutual action of bodies is compatible only with the finite universe, not with a quasi-Euclidean, infinite universe. According to Einstein's argument, an infinite universe is possible only if the mean density of matter in the universe vanishes, which is unlikely due to the fact that there is a positive mean density of matter in the universe \\footnote{However, in the appendix to the second edition of his book \\cite{ein22}, Einstein summarized Friedmann's world models and discussed a universe with vanishing spatial curvature and non-vanishing mean matter density, which differs from his original argument.}. In this paper, we propose the general world model for homogeneous and isotropic universe which supports Einstein's perspective on the physical universe. The outline of this paper is as follows. In Sec. \\ref{sec:metric}, we consider the effect of cosmic expansion on the physical space-time distance between neighboring events and describe how to define the line element for homogeneous and isotropic universes of various spatial curvature types. The metric and the cosmic evolution equations for flat, closed, and open universes are derived in Sec. \\ref{sec:nspace}. It will be shown that our universe is spatially closed. In Sec. \\ref{sec:some}, we reconstruct cosmic evolution histories of the closed world models, and derive interesting properties of the closed universe. In Sec. \\ref{sec:inflation}, we discuss whether the inflation theory is compatible with the closed world model or not. Conclusion follows in Sec. \\ref{sec:conc}. Throughout this paper, we adopt a sign convention $(+,-,-,-)$ for the metric tensor $g_{ik}$, and denote a 4-vector in space-time as $p^{i}$ ($i=0,1,2,3$) and a 3-vector in space as $p^{\\alpha}$ ($\\alpha=1,2,3$) or $\\mathbf{p}$. The Einstein's field equations are \\begin{equation} R_{ik} - \\frac{1}{2} g_{ik}R = 8\\pi G T_{ik}+\\Lambda g_{ik}, \\label{eq:einstein} \\end{equation} where $R_{ik}=R^{a}_{~iak}$ is the Ricci tensor, $R=R^i_{~i}$ the Ricci scalar, $T_{ik}$ the energy-momentum tensor, $G$ the Newton's gravitational constant, and $\\Lambda$ the cosmological constant. The Riemann curvature tensor is given by $R^{a}_{~ibk} =\\partial_b \\Gamma^{a}_{ki}-\\partial_{k}\\Gamma^{a}_{bi} +\\Gamma^{a}_{bn} \\Gamma^{n}_{ki}-\\Gamma^{a}_{kn}\\Gamma^{n}_{bi}$, with the Christoffel symbol $\\Gamma^{a}_{ik}={1\\over 2} g^{ab} (\\partial_i g_{kb}+\\partial_k g_{ib} -\\partial_b g_{ik})$. The energy-momentum tensor for perfect fluid is \\begin{equation} T_{ik} = (\\varepsilon_\\textrm{b} + P_\\textrm{b})u_i u_k - P_\\textrm{b} g_{ik}, \\label{eq:emtensor} \\end{equation} where $\\varepsilon_\\textrm{b}$ and $P_\\textrm{b}$ are background energy density and pressure of ordinary matter and radiation, and $u_{i}$ is the 4-velocity of a fundamental observer. We assume that the cosmological constant acts like a fluid with effective energy density $\\varepsilon_\\Lambda = \\Lambda/8\\pi G$ and pressure $P_\\Lambda = -\\varepsilon_\\Lambda$. The limiting speed in the special theory of relativity is set to unity ($c\\equiv 1$). ", "conclusions": "\\label{sec:conc} In this paper, the general world model for homogeneous and isotropic universe has been proposed. By introducing the world reference frame as a global and fiducial system of reference, we have defined the line element so that the effect of cosmic expansion on the physical space-time separation can be correctly included in the metric. With this framework, we have demonstrated theoretically that the flat universe is equivalent to the Minkowski space-time and that the universe with positive energy density is always spatially closed and finite. The open universe is unrealistic because it cannot accommodate positive energy density. Therefore, in the world of ordinary materials, only the spatially closed universe is possible to exist. The naturalness of the finite world with positive energy density comes from the Mach's principle that the motion of a mass particle depends on the mass distribution of the entire world. The principle is consistent only with the finite world because the dynamics of a reference frame cannot be defined in the infinite, empty world. The closed world model satisfies the Mach's principle and supports Einstein's perspective on the physical universe. We have reconstructed evolution histories of the closed world models that are consistent with the recent astronomical observations, based on the nearly flat FRW world models (Model I and II; Sec. \\ref{sec:some}). The present curvature radius of the universe is $a_0 = 25.7$ Gpc ($a_0=40.2$ Gpc) for Model I (Model II). The expansion histories of both models imply that the closed universe dominated by dark energy expands eternally. However, the currently favored flat FRW world exists only as a limiting case of the closed universe with infinite curvature radius that is expanding with the maximum speed ($\\dot{a}_0=1$, $\\ddot{a}_0=0$). From the local nature of the FRW metric (Sec. \\ref{sec:metric}) and of the proper time of a comoving observer (Sec. \\ref{sec:rel_friedmann}), it is clear that the FRW world model describes the local universe as observed by the comoving observer. Since the Newton's gravitation law can be derived from the Einstein's field equations in the weak field and the small velocity limits, the gravitational action at a distance usually holds at a local region of space on scales far smaller than the Hubble horizon size (e.g., \\cite{peebles80}). The proper Hubble radius $d_\\textrm{H}$ (Fig. \\ref{fig:horizon}, top) may provide a reasonable estimate of the characteristic distance scale where the Newton's gravity applies. The cosmic structures simulated by the Newton's gravity-based $N$-body method will significantly deviate from the real structures on scales comparable to $d_\\textrm{H}$. The variation of the comoving Hubble radius $\\chi_\\textrm{H} = d_\\textrm{H} /a$ also implies that in the past (future) the Newtonian dynamics was (will be) applicable on smaller region of space compared to the size of the universe (Fig. \\ref{fig:horizon}, bottom). In this paper, the history of the universe has been tentatively reconstructed based on cosmological parameters of non-flat FRW world models. The more general cosmological perturbation theory and parameter estimation are essential for accurate reconstruction of the cosmic history." }, "0710/0710.4914_arXiv.txt": { "abstract": "We present 2D simulations of the cooling of neutron stars with strong magnetic fields ($B \\geq 10^{13}$ G). We solve the diffusion equation in axial symmetry including the state of the art microphysics that controls the cooling such as slow/fast neutrino processes, superfluidity, as well as possible heating mechanisms. We study how the cooling curves depend on the the magnetic field strength and geometry. Special attention is given to discuss the influence of magnetic field decay. We show that Joule heating effects are very large and in some cases control the thermal evolution. We characterize the temperature anisotropy induced by the magnetic field for the early and late stages of the evolution of isolated neutron stars. ", "introduction": "The observed thermal emission of neutron stars (NSs) can provide information about the matter in their interior. Comparing the theoretical cooling curves with observational data\\cite{Yakovlev2004,Page2006} one can infer not only the physical conditions of the outer region (atmosphere) where the spectrum is formed but also of the poorly known interior (crust, core) where high densities are expected. There is increasing evidence that most of nearby NSs whose thermal emission is visible in the X-ray band have a non uniform temperature distribution\\cite{Zavlin2007,Haberl2007}~. There is a mismatch between the extrapolation to low energy of the fits to X-ray spectra, and the observed Rayleigh Jeans tail in the optical band ({\\it optical excess flux}), that cannot be addressed with a unique temperature (e.g. \\rxdieciocho\\cite{Pons2002}~, \\rbdoce\\cite{Schwope2007}~, and \\rxcerosiete\\cite{Perez2006}~). A non uniform temperature distribution may be produced not only in the low density regions\\cite{Greenstein1983}~, but also in intermediate density regions, such as the solid crust. Recently, it has been proposed that crustal confined magnetic fields with strengths larger than $10^{13}$ G could be responsible for the surface thermal anisotropy \\cite{Geppert2004,Azorin2006}~. In the crust, the magnetic field limits the movement of electrons (main responsible for the heat transport) in the direction perpendicular to the field and the thermal conductivity in this direction is highly suppressed, while remains almost unaffected along the field lines. Moreover, the observational fact that most thermally emitting isolated NSs have magnetic fields larger than $10^{13}$ G implies that a realistic cooling model must include magnetic field effects. In a recent work\\cite{Aguilera2007}~, first 2D simulations of the cooling of magnetized NSs have been presented. In particular, it has been stated that magnetic field decay, as a heat source, could strongly affect the thermal evolution and the observations should be reinterpreted in the light of these new results. We present the main conclusions of this work next. ", "conclusions": "The main result of this work is that, in magnetized NSs with $B> 10^{13}$~G, the decay of the magnetic field affects strongly their cooling. In particular, there is a huge effect of Joule heating on the thermal evolution. In NSs born as magnetars, this effect plays a key role in maintaining them warm for a long time. Moreover, it can also be important in high magnetic field radio pulsars and in radio--quiet isolated NSs. As a conclusion, the thermal and magnetic field evolution of a NS is at least a two parameter space (Fig~\\ref{fig_coupled}), and a first step towards a coupled magneto-thermal evolution has been given in this work. \\begin{figure}[htb] \\begin{center} \\psfig{file=BT_v1_1.ps,width=0.5\\textwidth,angle=-90} \\end{center} \\caption{Coupled magneto-thermal evolution of isolated neutron stars\\cite{Aguilera2007a}~: $T_s$ for the hot component as a function of $B$ and $t$. Observations: squares for magnetars ($B>10^{14}$~G), triangles for intermediate-field isolated NSs ($10^{13}$~G$